Of Syntliesis,: Co, Ni, Cu

Download as pdf or txt
Download as pdf or txt
You are on page 1of 248

Zeolite P,ncapsufatea of

Pe, Co, Ni, Cu antfcptf:


Syntliesis, Cliaracterization antf Catafysis
Jyothi Mariam John
Department of Applied Chemistry
Cochin of Science and Technology
Kochi. KERALA. 682 022
\Ian:h 201L"'
Certificate
This is to certify that the thesis entitled "Zeolite encapsulated complexes of Fe,
Co, Ni, Cu, and Pd: synthesis, characterization, and catalysis," submitted for the
award of the degree of Doctor of Philosophy of Cochin University of Science and
Technology, is a record of bona fide research work carried out by Ms. Jyothi
Mariam John under my supervision in the Department of Applied Chemistry.
Prof. K.K. Mohammed Yusuff
Department of Applied Chemistry
Cochin University of Science and Technology
Kochi-22
Preface
The growing demand for the design and development of environmentally benign
products and processes has led to the emergence of 'green chemistry.' One of the
basic thrusts of 'green chemistry' is on minimizing the production of waste in
chemical processes. In line with this need we undertook the studies on zeolite Y
encapsulated transition metal complexes and their application as catalysts for
some selective oxidation reactions of vital importance.
Encapsulation inside the zeolite cages makes the catalysts more stable.
Further, the framework prevents the complexes from dimerising. The topology of
the void influences the complex's geometry, and brings about changes in its
electronic and magnetic properties. The steric constraints in zeolite may distort
the encapsulated complex molecule to induce different stability and reactivity for
them. The ligands modify and fine-tune the electronic and steric environment of
the reaction site. Thus the zeolite encapsulated complexes become reaction
centers, which bind small molecules used for catalysis. The present thesis deals
with the synthesis, characterization, and catalytic activity studies of some zeolite
encapsulated complexes. The thesis is divided into nine chapters.
Chapter 1 is a general introduction on zeolite encapsulated complexes and
their application as catalysts in certain oxidation reactions, and Chapter 2
describes the materials and methods used in the present study. Chapter 3 is
divided into two sections. Section A deals with the synthesis and characterization
of Schiff bases of Fe(III), Co(I1), Ni(II), and Cu(U) complexes of the NN' -bis(3-
pyridylidene )-1 ,2-phenylenediamine and N.N' -bis(3-pyridylidene)-I,4-
phenylenediamine. Section B deals with Co(II), Ni(II), and CU(II) complexes of
salicylaldehyde semicarbazone encapsulated in the zeolite Y cage. Chapter 4 is a
comparative study on the neat and encapsulated complexes of
2-styrylbenzimidazole complexes of CU(II). Chapter 5 is about the zeolite
encapsulated dithiocarbamate complexes, and has two sections. Section A
presents the studies on Fe(III) dithiocarbamates, and Section B is on Co(III),
Ni(II), and CU(II) morpholinedithiocarbamates. Chapter 6 of the thesis deals with
the study of the zeolite encapsulated palladium complexes of DMG, SALSC, and
MOTe. Chapters 7, 8, and 9 focus on catalytic activity studies using the
synthesized encapsulated complexes. The general conclusions drawn from the
investigation are given at the end of the thesis.
Contents
Chapter 1 General Introduction
1
1. Introduction
2
1.1. Catalytic oxidation reactions
3
1.2. Zeolite encapsulated metal complexes
8
1.2.l. Zeolites
9
1.2.2. Classification of zeolites
11
1.2.3. Structural aspects of zeolites
11
1.2.4. Synthesis of encapsulated complexes
14
1.2.4.l. Flexible ligand method
14
1.2.4.2. Ship-in-a-bottle synthesis
15
1.2.4.3 . Zeolite synthesis method
17
1.2.5 . Characterization techniques
17
1.2.6. Zeolites as micro reactors
18
1.2.7. Guest host interactions between complex and support
19
1.2.8. Zeozymes: encapsulated complexes that mimic
21
enzymes
1.3. Scope of the present study
23
References 26
Chapter 2 Materials and methods
32
2.l. Introduction
33
2.2. Reagents
33
2.3. Synthesis of ligands
34
2.3.l. Preparation of dimethylglyoxime (DMG)
34
2.3.2. Synthesis of styrylbenzimidazole (SB)
34
2.3.3. Synthesis of carbodithioate and dithiocarbarnates
35
2.3.4. Synthesis of N,N'-bis(3-pyridylidene )-1,2-
35
phenelenediamine (SPO)
2.3.5. Synthesis of N,N'-bis(3-pyridylidene )-1,4-
35
phenylenediamine (SPP)
2.3.6. Synthesis of salicylaldehyde semicarbazone (SALSC)
36
2.4. Synthesis of zeolite Y encapsulated metal complexes
36
2.4.l. Preparation of sodium exchanged zeolite (NaY)
36
2.4.2. Preparation of transition metal ion exchanged zeolite
36
(MY)
2.4.3. Encapsulation of metal complexes in zeolite
37
2.4.3.l. Encapsulation by heating in a sealed ampule
37
2.4.3 .2. Encapsulation by refluxing the metal exchanged zeolite
38
with the ligand
2.5. S}nthesis of neat complex
38
2.6. Catalysis procedures
38
2.6.l. Decomposition of hydrogen peroxide
38
2.6.2. Hydroxylation of phenol to hydro quinone
39
2.6.3. Partial oxidation of benzyl alcohol to benzaldehyde
39
2.7. Characterization methods
39
2.7.l. Chemical analysis
40
2.7.1.1. Analysis of Si, AI, N a, and transition metal ion in the
40
zeolite sample
2.7.1.2. Detennination of metal in neat complex
41
2.7.1.3. Detennination of chlorine
41
2.7.2. CRN analysis
42
2.7.3. Surface area and pore volume analysis
42
2.7.4. X-ray diffraction studies
43
2.7.5. Spectroscopic analysis
44
2.7.5.1. Electronic spectra
44
2.7.5.2. Infrared spectra
45
2.7.5.3. EPRspectra
45
2.7.6. Magnetic measurement studies
46
2.7.7. Scanning electron microscopy
46
2.7.8. Thenno gravimetric analysis
47
2.7.9. Conductance measurement
47
2.8. Catalytic studies
47
2.8.1. Gas chromatography
47
2.8.2. Thin layer chromatography
48
References
49
Chapter 3 Zeolite-Y encapsulated complexes of the Schiff bases 50
N;N'-bis(3-pyridylidene )-1 ,2-phenylenediamine,
N;N'-bis(3-pyridyIidene)-I,4-phenylenediamine, and
salicylaldehyde semicarhazone
3. Introduction
51
Section A Zeolite encapsulated Fe(m), Co(D), Ni(D) and Cu(II)
52
N,N'-bis(3-pyridylidene)-I,2-phenylenediamine and
N,N'-bis(3-pyridylidene )-1 ,4-phenylenediamine
complexes
JA1. Introduction
52
3A2. Experimental
53
3A2.1. Synthesis of ion exchanged zeolites
53
3A.2.2. Synthesis of zeolite encapsulated Fe(III), Co(Il), Ni(II)
53
and Cu(II) complexes of N,N'-bis(3-pyridylidene )-1,2-
phenylenediamine, and N,N'-bis(3-pyridylidene )-1,4-
phenylenediamine
3A.3. Characterization techniques
54
3A4. Results and discussion
54
3A4.1. Ion exchanged zeolite
54
3A4.1.1. Chemical analysis
54
3A4.1.2. X - ray diffraction patterns
56
3A.4.l.3. Surface area and pore volume
58
3A.4.1.4. FTIR spectrum
58
3A.4.1.5. Diffuse reflectance spectra
60
3A.4.1.6. EPR spectrum
60
3A.4.2. Zeolite encapsulated complexes
62
3A.4.2.1. Chemical analysis
62
3A.4.2.2. Scanning electron micrographs
63
3A.4.2.3. X-ray diffraction patterns
63
3A.4.2.4. Surface area and pore volume
67
3A.4.2.5. Magnetic moment
68
3A.4.2.6. Diffuse reflectance spectra
70
3A.4.2.7. FTIR spectra
75
3A.4.2.8. EPRspectra
80
3A.4.2.9. TG analysis
84
Section B Zeolite encapsulated Co(JI), Ni(JI) and Cu(JI)
88
complexes of salicylaldehyde semicarbazone
38.1. Introduction
88
38.2. Experimental
89
38.2.1. Synthesis of zeolite encapsulated Co, Ni and Cu
89
complexes of salicylaldehyde semicarbazone
38.3. Characterization techniques
89
38.4. Results and discussion
89
38.4.l. Chemical analysis
89
38.4.2. X -ray diffraction studies
90
38.4.3 Surface area and pore volume
91
38.4.4. Magnetic moment
92
38.4.5. Diffuse reflectance spectra
93
38.4.6. FTIR spectra
96
38.4.7. EPRspectra
98
38.4.8. TG analysis
99
References 102
Chapter 4 A comparison of neat and encapsulated complexes of
104
copper 2-styrylbenzimidazole
4.1. Introduction
105
4.2. Experimental
105
4.2.l. Synthesis of neat Cu(SBhCh
106
4.2.2. Synthesis of zeolite encapsulated 2-styrylbenzimidazole
106
complex
4.3. Characterization techniques
106
4.4. Results and discussion
106
4.4.1. Chemical analysis
107
4.4.2. X-ray diffraction patterns
108
4.4.3 . Scanning electron microscopy
108
4.4.4. Surface area and pore \"olume
108
4.4.5. Magnetic moment
III
4.4.6. Diffuse reflectance spectra
III
4.4.7. FTIR spectra
113
4.4.8. EPRspectra
113
4.4.9. TG analysis
118
References
121
Chapter 5 Zeolite-Y encapsulated dithiocarbamates of Fe, Co,
123
Ni and Cu
5. Introduction 124
Section A Zeolite Y encapsulated Fe(III) dithiocarbamate and
125
carbodithioate complexes
SA. 1. Introduction
12S
SA.2. Experimental
126
SA.2.1. Synthesis of zeolite encapsulated iron carbodithioate
126
and dithiocarbamates
5A.3. Characterization techniques
126
5A.4. Results and discussion
127
SA.4.1. Chemical analysis
127
SA.4.2. X-ray diffraction patterns
127
SA.4.3. Scanning electron micrographs
128
5A.4.4. Surface area and pore volume
130
SA.4.S. Magnetic moment
130
SA.4.6. Diffuse reflectance spectra
131
SA.4.7. FTIR spectra
133
SA.4.8. TG analysis
138
Section B Zeolite encapsulated Co(m), Ni(II) and Cu(II)
139
morpholine-N-carbodithioate complexes
SB.l. Introduction
139
SB.2. E).']Jerimental
139
SB.2.1.
Synthesis of zeolite encapsulated Co(III), Ni(II) 139
and Cu(II) morpholine-N-carbodithioate complexes
SB.3. Characterization techniques 140
SB.4. Results and discussion
140
SB.4.1. Chemical analysis 140
5B.4.2. X-t:ay diffraction patterns 141
5B.4.3. Scanning electron micrographs
141
SB.4.4. Surface area and pore volume data
144
SB.4.5. Magnetic moment
144
5B.4.6. Diffuse reflectance spectra
145
SB.4.7. FTIR spectra
147
5B.4.8. EPR spectrum
150
5B.4.9. TG analysis
151
References
153
Chapter 6 Zeolite encapsulated palladium complexes
154
6.l. Introduction
155
6.2. Experimental
157
6.2.1. Synthesis of zeolite encapsulated palladium complexes
157
6.3. Characterization techniques
157
6.4. Results and discussion
158
6.4.1. Chemical analysis
158
6.4.2. X-ray diffraction patterns
158
6.4.3. Scanning electron microscopy
160
6.4.4. Surface area and pore volume
161
6.4.5. Magnetic moment
161
6.4.6. Diffuse reflectance spectra
162
6.4.7. FTIR spectra
164
6.4.8. TG analysis
169
References
171
Chapter 7 Catalytic activity ofzeolite encapsulated complexes 172
in the decomposition of hydrogen peroxide
7.l. Introduction 173
7.2. Experimental
173
7.2.1. Materials 173
7.2.2. Procedure for the decomposition of hydrogen peroxide
174
7.3. Results 174
7.3.1. Screening studies
174
7.3.2. Blank run
177
7.3.3. Factors influencing the decomposition of hydrogen
179
peroxide
7.3.3.1. Influence of the volume of hydrogen peroxide 179
7.3.3.2. Influence of the amount of catalyst 180
7.3.3.3. Influence of solvent polarity of the reaction solution
182
7.3.4. Addition ofpyridine
184
7.3.5. Recycled catalysts
185
7.4. Discussion
188
References
193
Chapter 8 Hydroxylation of phenol to hydroquinone using 194
zeolite encapsulated complexes
8.l. Introduction
195
8.2. Experimental
195
8.2.1. Materials
195
8.2.2. Procedure for the hydroxylation of phenol
196
8.3. Results
196
8.3.l. Screening studies
196
8.3.2. Blank run
198
8.3.3. Factors influencing the hydroxylation of phenol to form
198
hydroquinone
8.3.3.l. Influence of reaction time
198
8.3.3.2. Influence of the amount of catalyst
200
8.3.3.3. Influence ofH
2
0
2
: phenol (oxidantlsubstrate) ratio
202
8.3.3.4. Influence of solvents
203
8.3.3.5. Influence of reaction temperature
204
8.3.4. Addition of poison
205
8.3.5. Recycled catalyst
205
8.4. Discussion
206
References
213
Chapter 9 Catalytic activity of zeolite encapsulated complexes 214
in the partial oxidation of benzyl alcohol to
benzaldehyde
9.l. Introduction
215
9.2. Experimental
215
9.2.l. Materials
215
9.2.2. Procedure for the partial oxidation of benzyl alcohol to
216
benzaldehyde
9.3. Results 216
9.3.l. Screening studies
216
9.3.2. Blank run 218
9.3.3. Factors influencing the partial oxidation of benzyl 218
alcohol to benzaldehyde
9.3.3.l. Effect of reaction time
218
9.3.3.2. Effect of the amount of catalyst
220
9.3.3.3. Effect of oxidant: substrate ratio
221
9.3.3.4. Effect of solvents
222
9.3.3.5. Effect of reaction temperature
223
9.3.4. Addition of poison 224
9.3.5. Recycled catalyst
225
9.4. Discussion
226
References
230
Summary and conclusion 231
Chapter 1
General Introduction
1. Introduction
Modern man relies heavily on chemicals and has constantly been on the lookout
for new chemicals and easy means to obtain them. Catalysis, the method of
enhancing chemical processes by the use of appropriate material, has therefore
been an area of vital interest. Nearly 90 % of the products manufactured today
require processes which involve at least one catalytic step. As chemical processes
always hold a potential threat to the environment, the nature of the catalyst used
assumes special importance in relation to environmental protection. In the words
of R.A. Sheldon, " The chemical industry is under increasing pressure to minimize
or preferably eliminate waste production in chemical manufacture" [1]. The goal
now is to synthesize catalysts that do not get deactivated and can be recovered
with 100% efficiency.
The search for efficient catalytic processes having high activity and
selectivity is of perennial importance. The recent urge for environmental
protection has added a new dimension to the search for new catalysts. The need to
use regenerable and safe solid catalysts is being increasingly felt. Transition metal
complexes are accessible in a series of oxidation states and are good catalysts for
many reactions including oxidation. However, they have certain disadvantages
because of their homogeneous nature. Furthermore, many of the transition metal
complexes are expensive to purchase and prepare. One method to overcome these
limitations is to encapsulate them in zeolite pores. Such immobilisation of the
complexes on to a support facilitates simple extraction as well as recycling,
ensuring commercial advantages besides ease of manipulation. Efforts are being
made by investigators worldwide to design and develop zeolite catalysts for active
and selective organic transformations, especially for refining organic chemicals
derived from the oxidation of hydrocarbons [2,3], and most of them use catalytic
oxidations.
2
Chemical processes involving catalysis are often preferred over their non-
catalytic alternatives as they take place more efficiently under milder conditions,
producing higher yields with lower energy consumption, and minimizing
environmental pollution. Further introduction of new and safe processes leads to
high quality products. Active research is being carried out in the following aspects
to develop new catalytic systems:
spectroscopy of catalysts and catalyst models
kinetics of catalytic reactions
quantum chemical calculations for reactants, intermediates and products
from measured spectra and quantum chemical calculations
microkinetic modeling.
The transformation of functional groups by selective oxidation using metal
complexes as catalysts is a basic reaction in organic chemistry. Oxidation reactions
are the preferred processes in chemical synthesis, because of the ease with which
the primary oxo product can be converted into secondary products like acids, diols,
amines or esters [4]. By selecting the appropriate catalyst and reaction conditions,
it is possible to direct the reaction along one selected pathway (lower energy
pathway) to obtain the desired product.
1.1. Catalytic oxidation reactions
The ability of transition metals to exist in different oxidation states makes
them excellent catalysts for many reactions. Several transition metal complexes
act as catalysts for oxidation reactions. Molecular oxygen and air are attractive
oxidants as they are inexpensive and yield no environmentally harmful side
products. But since catalytic processes with these oxidants require elevated
temperatures and pressures, oxidation is generally carried out using active oxygen
reagents such as peroxides. Among the peroxides, hydrogen peroxide is one of the
most readily available peroxide [5,6].
Even though there are many types of useful oxidation reactions, the present
work is limited to the study of three oxidation reactions, namely, the
decomposition of hydrogen peroxide, hydroxylation of phenol, and oxidation of
benzyl alcohol. Therefore a brief review on these oxidation reactions is presented
in this chapter.
Hydrogen peroxide is industrially and chemically important because of its
use in a variety of processes [7]. The decomposition of hydrogen peroxide is
considered as a standard reaction for the determination of the catalytic activity of
metal complexes [8,9]. The reaction can be catalyzed by several transition metal
ions and transition metal complexes, provided the coordination sphere is accessible
to H202 or HOO- ion [10,11]. The disproportionation of hydrogen peroxide
catalyzed by metal complex has received much attention because this reaction is
catalyzed by the naturally occurring heme enzyme catalase [12]. This enzyme is
used by some bacteria to catalyze the conversion of hydrogen peroxide to oxygen
and water. As hydrogen peroxide is an unwanted byproduct in many redox
reactions involving oxygen in living systems, this reaction, an effective method for
removing hydrogen peroxide, continues to receive the attention of bioinorganic
chemists.
Sharma and Schubert have made a detailed study on the kinetics and
decomposition of hydrogen peroxide by copper complexes and chelates [13,14].
They have observed the effect of the chelate ring size on hydrogen peroxide
decomposition [15]. Efforts have been made to prepare coordination compounds
containing dioxygen groups and subsequently to use these complexes for catalytic
oxidation reactions. Examples of such systems can be found in nature. SALEN
N,N'-ethylenebis(salicylideneamine)] complexes and SALSC (salicylaldehyde
semicarbazone) complexes of transition metals have been reported to be mimics of
catalase [16] and superoxidedismutase [17] which are known to catalyze the
decomposition of hydrogen peroxide and superoxide ions. One of the major
4
disadvantages of metal SALENS in homogeneous solution is the formation of 11-
oxo dimer and other polymeric species leading to catalytic deactivation.
All the above-mentioned oxidation reactions are catalyzed by
homogeneous catalysts. Homogeneous systems have several drawbacks; these
catalysts show a tendency for self-degradation, and there is difficulty in reusing
them. Thus homogeneous catalysts are found to have limited commercial
application in large-scale preparations, as the homogeneous systems lack
continuous processing technology and recycling ability. The expense for catalyst
recovery is also high. When the products are not readily separated from the
catalyst, homogeneous catalysts can cause product contamination and catalyst loss.
For these reasons many homogeneous processes are not used on an industrial scale
in spite of their advantages.
The other oxidation reaction selected for study is phenol hydroxylation.
Phenol, a by-product in many industries, is used as a precursor of many organic
petrochemical compounds. It is readily biodegraded in the neutral environment if
its concentration is low, but artificial methods have to be used for its degradation if
the concentration is high. The accumulation of waste containing phenol leads to
serious environmental problems. Researchers are now trying to develop efficient
and inexpensive processes for the disposal of this type of waste. Highly specific
first step oxidation products are obtained with mild oxidizing agents and suitable
catalysts, while quinones and other degradation products are obtained in the
presence of strong oxidants.
One method for the disposal of phenol is its conversion to industrially
useful diphenols (catechol and hydro quinone). But vigorous reaction conditions
like high temperatures and pressure are required for the processes and make such
processes uneconomical for phenol oxidation [18]. Further industrial applications
of diphenols and their derivatives motivated the discovery of new methods for
their production from phenol. Catechol and hydroquinone belong to the class of
5
low volume but high value fine chemicals. The principal large-scale application of
hydroquinone is in photography, as a photographic film developer [19].
Phenol hydroxylation has been carried out using different oxidants and a
variety of processes. The processes using hydrogen peroxide as the oxidant have
disadvantages owing to its decomposition during reaction and the threat of
explosion, and it is not economically favorable in many oxidation processes. But it
is still preferred because of the environmental advantage. The most efficient
contemporary processes involve the oxidation of phenol with H
2
0
2
to produce a
mixture of catechol and hydro quinone. Leading producers following this route are
Rhone - Poulenc, Brichima,and Enichem.
The Rhone-Poulenc process [20] is based on phenol hydroxylation by
hydrogen peroxide catalyzed by strong mineral acid. In the industrial process,
perchloric acid is used in its azeotropic mixture with water. The reaction is
performed at 90C using phenol, H202 and HCI0
4
in the molar ratio 20:1 :0.1. The
disadvantage of the process is the formation of a high amount of tar.
Hydroxylation of phenol using hydrogen peroxide and iron based Fentons reagent
was the procedure adopted by Brichima in Italy [21-25]. During the reaction
ferrous salts are consumed, breaking off the redox reaction [26]. Ferrocene also
has good catalytic properties for this reaction [27]. The formation of high amounts
oftar is the major disadvantage of this method also.
Hydroxylation of phenol is also catalyzed by simple metal ions and their
complexes [28,29]. In spite of the high activity of these catalysts, they have the
inherent disadvantages of homogeneous catalysts. A notable disadvantage of
homogeneous catalysis is the high ortho : para ratio which is close to 2.
The conventional methods of phenol hydroxylation using heterogeneous
catalysts result in a mixture of ortho and para hydroxy phenols. Many of the acid
6
catalyzed reactions which require very corrosive acids in homogeneous catalysis
can be carried out using zeolites and related acid catalysts [30,31].
Zeolite-based catalysts are now used for the production of hydroquinone as
they offer advantages over existing processes. The shape selective nature of
zeolites favours the production of hydroquinone [32,33]. The hydroxylation of
phenol with hydrogen peroxide has been carried out in the presence of zeolite
modified with rare earth catalyst [34]. Hydroquinone formation takes place only in
the intra crystalline space of the zeolite catalyst whereas catechol and tarry
products can be formed either on the outer surface of the catalyst or on the inner
surface when sufficient space is available [35,36]. The first industrial application
ofa zeolite catalyst for phenol hydroxylation was in the Enichem process [37,38].
Partial oxidation of benzyl alcohol involving the metal complexes was also
studied as a part of the work embodied in the present thesis. This reaction is of
great importance in the preparation of fine chemicals, as it makes the production of
intermediate compounds easy. Side chain oxidation reactions resulting in an
intermediate product with high selectivity is not very common with neat
complexes. There are reports of studies on benzyl alcohol oxidation to
benzaldehyde taking place using atmospheric oxygen and hydrogen peroxide. On
adding hydrogen peroxide to metal complexes the corresponding per-oxo
compounds are generated. The per-oxo compounds formed provide an attractive
route for the preparation of synthetic intermediates and other oxygen containing
organic substrates. Metal complexes have thus attracted attention as dioxygen
activating catalysts [39-41]. Benzyl alcohol to benzaldehyde oxidation has been
carried out using hydrogen peroxide as oxidant and eo(n) acetate as catalyst at 70
c [42]. Sumathi has reported benzyl alcohol oxidation using perovskite catalysts
[43]. A convenient and efficient method for the synthesis of benzaldehyde from
benzyl alcohol using ruthenium oxo-complexes has been reported by Seok [44].
1.2. Zeolite encapsulated metal complexes
In catalysis the pnmary goal is to promote reactions that have high
selectivity with high yields. Site isolation of the active centres by bonding to the
support is a convenient method to achieve high selectivity. This procedure
stabilizes the complex species and makes them act as heterogeneous solid catalysts
for liquid phase oxidation [45].
There are several methods of supporting homogeneous catalysts, and such
supported catalysts can be classified into the following categories:
1. supported liquid phase catalysts
11. catalysts anchored on functionalised solids
111. polymerized complexes
IV. tethered or grafted complexes
v. intercalated catalysts, and
VI. encapsulated catalysts.
Of these, the present study employs the method of encapsulation. Many desired
effects in catalysis that can be realized neither by homogeneous systems nor by
conventional heterogeneous systems can be achieved by encapsulating complexes
in zeolite pores. Heterogenised homogeneous catalysis has emerged as a method
having the advantages of both homogeneous catalysis (specificity and activity) and
heterogeneous catalysis (easy removal of catalyst) [46]. Such catalysts are
regarded as third generation catalysts.
Metal complexes immobilized in zeolite super cages are found to have
almost the same activity or more as compared to free complexes in solution.
Because of size restriction, the complex stays in the zeolite and cannot diffuse out.
This provides a route for dispersing the complexes in the solid structure. The main
objective is to improve the stability of the complex by reducing dimerisation or
aggregation of the complex. Zeolite encapsulated complexes are now commonly
used as active catalysts in a wide range of chemical reactions [47-77].
8
1.2.1. Zeolites
Discovered by the Swedish mineralogists Cronstedt in 1756 [78] zeolite
crystals contain water, which was found to be lost when heated. This led Cronstedt
to call the mineral 'zeolite,' which in Greek means 'boiling stone.' There is
another story, which is contrary to the popular belief that he named them after the
Greek words meaning boiling stone. According to this story Zeolite was named
after Cronstedt's dog that dug up the rocks when it went out for a walk. Zeolite is a
Swedish word meaning 'dog.' The story goes on to say that the first zeolite
identified was called 'stilbite,' because his dog was still biting it when he got
home. Later it was noted that zeolites showed the properties of reversible water
loss [79]. Chabazite was found to have the ability to ion exchange and
differentially absorb molecules. Several workers in the 1800's noted the ability of
zeolite to remove odour from air [80]. During the period 1930 - 32 Mc Bain in his
studies on zeolites observed their capacity to absorb small molecules from large
ones, just like a sieve. Hence the name 'molecular sieves' was given to zeolites.
This sieving activity is possible because zeolites are micro porous, highly
crystalline substances. These zeolites have an internal structure that can be easily
tailored to absorb any number of species.
The industrial use of zeolites as catalysts and molecular sieves started after
Barrer's pioneering work in the early 1940's. This prompted chemists to seek
synthetic methods for the preparation of pure zeolites. Ever since, many synthetic
species have been reported, and most of them are without any natural counterparts.
More than 150 framework structures have been reported so far. These micro
porous solids are found to occupy a prominent place in the landscape of solid state
and material science, heterogeneous catalysis and clean technology. The first
reported synthetic zeolite was Linde A. This was synthesized in 1949 by Milton,
working at Union Carbide. This was found to have the formula
M xlm [AI xSi (I-x) O
2
]. nH
2
0
where M is a cation of valency m.
In the 1950's it was observed that zeolite based catalysts, especially those
containing rare earth ions and protons, possessed activity much greater than that of
conventional silica - alumina catalysts for cracking hydrocarbons. The economic
success of synthetic zeolites, particularly as catalysts, prompted scientists to seek
alternative sources of their natural counterparts. The shape selectivity of the
materials further enhanced their chemical usefulness. Great efforts were made to
find new molecular sieves, with tunable properties, altered / tailored structures, or
chemical composition. Crystalline faujasites were synthesized in the 1960's. The
synthesis of the pentasil family ofzeolites, ZSM-5, and ZSM-ll was also reported.
These zeolites function in a shape selective manner in certain alkylation,
isomerization, and dehydrative syntheses as well as in a variety of delicate
molecular separations [81].
In the early 1980's there emerged a new, large, family of micro porous
aluminophosphates (ALPO's) [82]. There are many ALPO's or MAPO's (frame
work substituted metal alumino phosphates where M could be Mg, Mn, Co, Fe,
etc.) that exhibit framework structures that are the same as those of either natural
or synthetic zeolites. In the early 1980's Italian workers discovered that Ti (IV)
ions isomorphously substituted for Si (IV) in open framework, yielded T S-1 and T
S-2 zeolites [81]. They show high activity and selectivity in oxidation reactions.
Another novel open structure titanosilicate ETS-I0 was synthesized later [83].
Isomorphous substitution of Si
4
+ by AI
3
+ or Ga
3
+ is an important way to modify
zeolite properties and is of much interest in zeolite chemistry [84].
The recent productions of mesoporous siliceous solids have opened up a
new era to pure and applied scientists. MCM-41, a large-pore mesoporous
structure was reported by Kresge et al. [85] and Beck et al. [86]. When these are
modified by replacing Si with AI or B they act as ideal shape selective catalysts for
bulky molecules. With the high surface area and thermal stability of me so porous
10
supports, opportunities exist not only for heterogenising homogeneous catalysts but
also for assembling biomimetic catalysts.
1.2.2. Classification of zeolites
Zeolites are classified in different ways. They are sometimes classified on
the basis of Sil Al ratio. Typical examples of low Sil AI ratio zeolites are A and X,
which have Si/AI ratio between 1 and l.5. Intermediate zeolites Y, L, and erionite
have Sil Al ratio between 2 and 5. High Sil AI ratio is seen in ZSM-5, ZSM-11,
EU-2, mordenite, ferrierite etc.
Barrer has classified zeolites into three groups based on the difference in
their effective pore diameter [87,88]. They are small pore zeolites (eg. Linde A,
Chabazite), intermediate pore zeolites (eg. ZSM-5, ZSM-ll, EU-2, ferrierite etc.)
and large pore zeolites (eg. Linde-L, Zeolite X and V). Mortier's classification of
zeolites is based on secondary building unit (SBU) [89].
1.2.3. Structural aspects of zeolites
Zeolites are members of the tectosilicate family of minerals [84,90-92].
They are highly crystalline, perfect micro porous materials. Structurally the zeolite
framework consists of Si0
4
and Al0
4
tetrahedra (T0
4
tetrahedra, where T = Si/AI)
joined by shared oxygen bridges forming three-dimensional network. The corner
linkages are accomplished through T -0-T bridges. The tetrahedra can arrange
themselves in many ways to form different porous crystal structures.Each oxygen
is shared between two neighbouring tetrahedra. The tetrahedra (primary building
unit) form rings of various sizes, which are linked to form complex units
(secondary building units) [93]. Twenty-four tetrahedra (silica and alumina) form a
sodalite unit, which is a cubo-octahedron with 24 vertices and consisting of six
four member rings (or four faces) and eight six member rings (or six faces). The
11
sodalite units formed in this way are the basic building units of X and Y. The
sodalite unit encloses a cavity known as sodalite cavity or cage.
Zeolite X and Y are formed by joining of the hexagonal 6 - membered
faces of sodalite unit (through D6R). The sodalite units are arranged like the
carbon in diamond and generate a large cavity or super cage with 12 -, six and four
- faces. Access to the super cage (a. - cage) is restricted by a 12 - face or window
[94]. The a. cage has a diameter of 13 A and cage mouth opening of about 7 - 8 A,
while the cages have a diameter of 6.6 A and an aperture of 2.2 A. In the case of
zeolite Y, 4 small windows open into the 13 A super cage. The Si/AI ratio of
zeolite X is - 1.25 while that of zeolite Y is - 2.5. The unit cell formula and other
characteristics of X and Y zeolites are given in Table 1.1.
Silicon can take four oxygen atoms around it in a tetrahedral combination
maintaining electrical neutrality. When a trivalent aluminium atom is similarly
bound to 4 oxygen atoms, the resultant aluminium tetrahedron will have a negative
valency. The net negative charge on the lattice is counterbalanced by exchangeable
cation at the following well-defined sites [95]: centre ofD6R (Site I), inside of -
cage adjacent to D6R (Site I), near the single - ring, but outside the sodalite in the
large cavity (Site II), inside the sodalite cage at the centre of single-6-ring (Site
n), and the wall of the large cavity (12 - ring aperture) (Site ill). The presence of
aluminium in the zeolite framework has several effects like introduction of an
electrical field (this makes the zeolite more hydrophilic), introduction of catalytic
centers, introduction of ion-exchange properties and lowering the thermal stability
of the zeolite.
Crystalline aluminosilicates are thermally and hydrothermally stable
because of the structure resulting from the interconnection of Si and AI tetrahedra
through the oxygen bridges. The stability increases as Si/AI ratio in the structure
increases. The zeolite structure remains stable at temperatures as high as 700C
[96].
12
Figure 1 The cavity and channel structure of zeolite Y
13
Table 1.1 Properties of zeolites X and Y
Mole ratio a cage ~ cage
~ ~ ~ ~ - - - . - - - - - - - . ... "."" .................................... " .....................
Zeolite Unit cell formula
Si0
2
/Ah 0
3
Diameter Aperture
(A) (A)
X N a86 (AI0
2
)86 (Si0
2
)106 nH
2
0 25:1 13 8
Y N a56 (AI0
2
)56 (Si0
2
)136 nH
2
0 5: 1 13 8
This structural diversity of the zeolites discussed above is responsible for the wide
range of interesting zeolite properties such as ion exchange capacity, specific
absorption behaviour, catalytic activity due to acidity, shape selectivity caused by
the size and polarity of molecules, thermal stability, and wide flexibility for
adjustment by isomorphous substitution of framework constituents. However,
these porous materials present some specific limitations in catalytic applications,
such as pore plugging, poisoning, migration, leaching, and structural defects
involving guests in general.
1.2.4. Synthesis of encapsulated complexes
Several methods have been used for the preparation of transition metal
complexes in the zeolite matrix.
1.2.4.1. Flexible ligand method
This idea was first introduced by Herron to prepare Co-SALEN complexes
in the supercages of faujasite [97]. This method takes advantage of the flexibility
of the ligand which enables it to enter the cavities of the zeolite host material
through the restricting windows. The SALEN is able to enter the pores of zeolite
14
due to possible free rotation around the carbon-carbon sigma bond which connects
two salicylidene moieties of the ligand. The ligand gets access to larger cages
where they react with the transition metal ions previously introduced into the
lattice. Once the ligand has entered the zeolite cage and complexed with
previously exchanged transition metal ions, the complex adopts a square planar
configuration, which is unable to escape from the zeolite void. Later on SALEN
complexes with Mn(II) [98], Fe(III) [99], Rh(II) [100], and Pd(II) [101] were
synthesized within zeolite cages.
Research in this field continued with variations in transition metal ion as
well as the structure of the Schiff base ligands [98,102]. The resulting compounds
have been shown to act as oxygen carriers mimicking hemoglobin and also as
selective oxidation catalysts [97,103]. This same synthetic method was used for
the preparation of complexes of transition metal ions with bipyridine. By adjusting
the ligand to metal ion ratio and the temperature of synthesis highly selective bis-
or tris- coordination complexes were synthesized [104,105]. These encapsulated
complexes were found to exhibit good catalytic activity [106]. Cr(III), Fe(III),
Bi(III), Ni(II), and Zn(II), complexes of N,N'-bis(salicylidene)propane-l,3-diamine
encapsulated in zeolite Y were also prepared by the flexible ligand method. The
encapsulated materials were found to be active catalysttfor the decomposition of
H
2
0
2
[69].
1.2.4.2. Ship-in-a-bottle synthesis
The ship-in-a-bottle method was first suggested by Romanovsky and
coworkers to encapsulate transition metal phthalocyanines in zeolites [107-109].
The first stage in the synthesis is the introduction of metal via ion exchange pre-
adsorption of a labile metal complex (such as a carbonyl or a metallocene)
followed by reaction with 1,2-dicyanobenzene. This procedure was adopted for the
synthesis of phthalocyanine complexes of cobalt, nickel, copper, iron, and
manganese [110-116].
15
This complex was also prepared by heating ion exchanged zeolite with 1,2-
dicyanobenzene (DCB) in vacuum at temperature between 250 DC and 350 DC
[I1O]. Complexation takes place according to the stoichiometry given by the
following equation
M2+y+ 4DCB + H
2
0 MPcY + O
2
+ 2H+
The optimum temperature varies with the type of transition metal ion and the
structure of the zeolite. The optimum complexation temperature can be determined
by differential scanning calorimetry [111,113]. The ion exchanged zeolite used for
this was not strongly dehydrated as water acts as an electron source for this
process [110].
Another method of ship-in- a- bottle synthesis involves the introduction of
the transition metal in zeolite by the absorption of the corresponding carbonyl
complex in the zeolite, followed by the synthesis of the phthalocyanine ligand
around the transition metal. This can happen in two ways. The carbonyl is first
decomposed thermally or photochemically whereby small metal clusters are
formed, or a direct ligand exchange of CO by 1,2-dicyanobenzene occurs at low
temperatures. The carbonyl route for the reaction is represented by the equation
MPcY+nCO
This indicates that no acid sites (no protons) are produced during complex
formation. The reaction can be carried out using metallocenes which act as
transition metal supplying agents. Here again small amount of water is required for
phthalocyanine synthesis.
M(Cp)z + Y + 4DCB + H
2
0 MPcY + 2Cpen + O
2
An advantage of this method is that the zeolite is free from uncomplexed metal
speCIes.
Zeolite encapsulated substituted phthalocyanines were also synthesized.
The syntheses of intrazeolite tetra-t-butyl-substituted iron-phthalocyanines, and
perfluorophthalocyanines of iron, cobalt, copper and manganese have been
reported [117-120].
16
The method was used for the preparation of porphyrin type ligands in the
supercages of zeolite Y. Nakamura et al. have reported the encapsulation of iron-
and manganese- tetramethylporphyrins in zeolite Y [121]. Encapsulation was
carried out by refluxing transition metal exchanged zeolite with pyrrole and
acetaldehyde in methanol solution. This method was also used for the synthesis of
tetraphenylporphyrins from pyrrole and benzaldehyde [122]. The same method has
been used for the synthesis of zeolite encapsulated complexes of rhodium.
[123,124].
1.2.4.3. Zeolite synthesis method
Zeolite synthesis method is one of the recent methods for synthesizing
zeolite encapsulated complexes. In this method, a zeolite structure is synthesized
around the pre-formed complex [125]. The transition metal complex has to satisfy
the following conditions for this type of synthesis:
the complex should be stable at the temperature and in the
hydrothermal and pH conditions under which zeolite synthesis
takes place
it should be sufficiently soluble in the synthesis medium to enable
random distribution of the complex in the synthesis mixture.
The zeolite synthesis method is advantageous in that it reqUlres only mild
preparation conditions and it yields well-defined encapsulated metal complexes.
1.2.5. Characterization techniques
A variety of experimental techniques have been developed for the
characterization of the catalysts and for tracing the course of catalytic reactions.
Characterization of zeolite samples is carried out to get a better understanding of
the acid sites, distribution of complex in the host, the degree of complexation, the
17
influence of host guest interaction on the structure of the complex, the stability of
the encapsulated complex compared to that of the neat complex in the solid or in
the neat dissolved state, the ability to exhibit free coordination sites for catalysis,
absence of surface species, and preservation of zeolite framework even after
encapsulation. The techniques used for the characterization of zeolite encapsulated
complexes are: chemical analysis, thermal analysis (TG, DTG, DT A), acidity
measurements by temperature programmed desorption, x-ray diffraction methods,
XPS, ESCA, SEM, TEM, ESR, infrared and Raman spectroscopy, N1'v1R. and
magic angle spinning nuclear magnetic resonance, Mossbauer spectroscopy,
diffuse reflectance spectra, magnetic susceptibility measurement, surface area and
pore volume analysis, electrochemical methods (cyclic voltarnmetry), molecular
modelling and simulation. The characterization studies of zeolite encapsulated
complexes have been reviewed by Ekloff and Emst [126].
1.2.6 Zeolites as micro reactors
An elementary molecular catalytic reactor refers to organized and
constrained media. which provide cavities or/and surfaces to accommodate the
substrate molecules and allow chemical reactions to occur [127]. It may contain
components. which generate a field of molecular interactions, securing the
molecular recognition, deciding the specificity of the reaction, and be sufficiently
mobile to adapt itself to conditions for a proper mutual orientation of reactants
determining the regio- and stereo- selectivity [128].
Like a molecular catalytic reactor the zeolite encapsulated complexes
contain active centers composed of one or more transition metal atoms surrounded
by appropriate ligands with one or more vacant coordination sites. The intrinsic
properties of the active center depend on the number and type of metal atoms, their
orbital structure and occupancy, number of vacant coordination sites, and the
properties of the ligand. Its extrinsic properties are functions of the type of the
support, its type of bonds, its electronic structure as well as structure of the
18
surface. The net negative charge of the zeolite framework and the distribution of
the positive charges of the cations can produce an electric field, which might
modify the properties of the molecule in the cage as in a molecular catalytic
reactor.
1.2.7. Guest host interactions between complex and support
The influence of zeolite moiety on the structure and reactivity of the
encapsulated complex has to be understood to facilitate its use as catalysts in
chemical reactions. According to Mochida et al. an isolated metal ion attached to a
molecular sieve would be expected to behave quite differently from a metal ion in
an ordinary metal oxide [129]. Zeolites with low Si/AI ratios have a great number
of compensating cations which will produce high electrostatic fields and field
gradients in channels and cavities. The electric field produced could reach up to
6-7 V at distances of a few angstroms from the cation site. Fields of this magnitude
in the internal voids of the pores stabilize positive species in a similar way as
highly polar, non-nucleophilic solvents do [130]. This is the reason why a variety
of organic cations can be easily generated and stabilized inside zeolite [131,132].
The presence of ligands attached to the metal ion can influence the
selectivity of metal exchanged zeolites. The topology of the void has an influence
on the complex's geometry and can bring about changes in its electronic and
magnetic properties. The steric constraints in zeolite may distort the encapsulated
complex molecule to induce different stability and reactivity for them. Spectral
and electrochemical studies were carried out study the host guest interaction in
zeolite encapsulated complexes. These interactions would enhance the catalysis of
molecules that are in zeolite pores [133,134].
Dimeric copper acetate and copper chloro acetate encapsulated in zeolite Y
catalyse the o-hydroxylation of phenol to catechol. The turn-over frequency for
phenol conversion increases upon encapsulation. Due to enhanced Cu-Cu binding
19
upon encapsulation, the strength and lability of Cu-phenolate and Cu-dioxygen
bonds are modified by a trans axial ligand effect, which accounts for the enhanced
reactivity of the encapsulated complex [135].
Zeolite encapsulated Co(II) , Ni(II) , and CU(II) complexes of 3-
formylsalicylic acid were synthesized and used for the catalytic oxidation of
benzyl alcohol. Of the three, the CU(II) complex has remarkable efficiency and is
stable enough to be recycled without much deterioration [136].
Comparative study of neat cobalt SALEN complex and the encapsulated
cobalt SALEN complex was carried out and marked difference in behaviour was
noted. Cobalt SALEN complex within the a. cage identified by the reflectance
spectrum is almost identical with that in solution. The pyridine adduct of the
encapsulated complex shows an affinity for oxygen [97]. The EPR spectrum of
this corresponds to the Co(SALEN)(Py)(02) complex. The formation of
peroxodimer in neat complex deactivates it. This is prevented in the zeolite matrix
and the oxygen adduct can be dissociated by simple evacuation [137].
Chandra and Ratnasamy have reported the decomposition of hydrogen
peroxide using substituted copper SALENS encapsulated in zeolite. The rate of
decomposition on using substituted SALENS has been found to come close to that
of enzyme catalase [16]. There are further reports by Ratnasamy et al. that
encapsulated SALEN compounds act like catalase [138]. Cu-SALEN was used for
the decomposition of hydrogen peroxide as well as for the oxidation of phenol and
p-xylene [139]. Another recent report is by Maurya et al. [69].
Cyclic voltammetric experiments with encapsulated complexes provide
information concerning the redox couple near the surface of the zeolite. This
highlights the importance of combining electrochemical methods with
conventional spectroscopic techniques (IR, UV-VIS and EPR) to assess the
differing nature and distribution of the entrapped transition metal complexes [140].
20
Herron et al. have reported the changes in the properties of nickel carbonyl
on encapsulation [141]. Encapsulated species has the formula Ni(CO)3 whereas no
such reactive species can exist in solution phase since it is highly reactive. These
examples indicate the influence of the zeolite matrix on the stability of the
complex.
1.2.8. Zeozymes: encapsulated complexes that mimic enzymes
Enzymes are powerful catalysts in life processes, and are responsible for
many complicated chemical reactions in living systems, which would not take
place in their absence. An enzyme is composed of a metal complex which acts as
the active site embedded in a large protein structure. The shape assumed by the
protein knots is determined by the sequence of amino acids making up the protein.
The protein part protects active sites from self- destruction and helps to induce
selectivity for a particular product. It exhibits high chemo-, regio-, and stereo-
selectivities for the specific transformation to take place. There exists a
complementary relationship between the active site of the enzyme and the
substrate. Zeolite encapsulated complexes exhibit this characteristic feature of
enzymes, and so they are called zeozymes. Thus, zeozymes are 'mimics' of
enzymes in zeolites [53]. Metal complexes encapsulated inside zeolite cage exhibit
enzyme-like activity, the zeolite part protecting the complex and ensuring proper
steric environment for the complex and the reaction pathway. The channels and
cages in the zeolite framework are similar to those created by the protein structure
of natural enzymes [142-144].
Enzyme mimicking involves the building of active centre of enzyme into
the matrix which allows a larger operational temperature domain and broader
spectrum of solvents. This is because the operational domain of enzyme is narrow
as far as temperature and solvents are concerned. This permits the immobilized
active part to be used like heterogeneous catalysts in reactors in which the liquid or
21
gaseous substrate is passed over a solid material. Non-natural metallo-enzymes are
studied because of the non-natural activity that may be expected from them. The
advantage of such techniques lies in the ease with which the catalyst can be
separated from the reactants.
Zeolite encapsulated porphyrin type complexes mimicking enzyme
cytochromeP-450 have received the attention of researchers [121,122,145]. Zeolite
encapsulated tetraphenyl porphyrins have been synthesized from pyrrole and
benzaldehyde. This complex mimics cytochromeP-450, which enhances the
selective oxidation of saturated hydrocarbons using hydrogen peroxide as oxidant
[136,137]. Several mimics of cyto chromeP-45 0 have been reported [142,146,147].
According to these reports encapsulation prevents the metal complexes from
dimerising. Further, the internal channels impose selectivity in the same way as do
the channels created by the protein structure of an enzyme.
Another efficient mimic of cytochromeP-450 is iron-phthalocyanine
complex encapsulated in zeolite Y which in turn IS embedded m
polydimethylsiloxane (PDMS) membrane [53]. This is the first zeozyme to be
incorporated in PDMS [148,149]. When this catalyst was brought in a membrane
set up for the oxidation of cyclohexane with TBHP, a remarkable six fold increase
in activity was observed as compared with the best possible experimental set up
for nonembedded FePcY. The reason for this effect was the influence ofPDMS on
the relative amounts of reagents that were sorbed in the zeolite pore. This system is
environmentally friendly. When Zn-phthalocyanine zeolite Y incorporated in
PDMS was used in the photosensitized oxidation of 1-methyl-1-cyclohexene in
ethanol, very high activity was observed [150]. Another zeozyme [Mn(bpyhf+y
was incorporated in PDMS for the epoxidation of olefines. Good catalytic activity
was obtained for reagents that diffused easily through the polymer and sorbed well
in the zeolite [151]. Zeolite encapsulated Co SALEN is also suggested as synthetic
mimic for studying the oxygen binding properties of haemoglobin [97,143]. These
adducts show excellent resistance to auto oxidation even at elevated temperatures.
22
In contrast to haemoglobin, these complexes exhibit negative co-operativity
between the cobalt binding sites.
The oxygenase mimicking activity of copper acetate dimers in the
regioselective ortho-hydroxylation of L-tyrosine to L-dopa is enhanced on
encapsulation in zeolite Y [152]. The spectra of this zeozyme reveals that the Cu-
eu separation in the dimer shortens to 2.4 A in the encapsulated state from 2.64 A
in the neat complex, providing greater overlap of metal orbitals in the encapsulated
stage [137]. The complex catalyzes the o-hydroxylation of phenol to catechol. Due
to enhanced Cu-Cu binding upon encapsulation, the strength and lability of the Cu-
phenolate and Cu-dioxygen bonds are modified by a trans-axial ligand effect. This
explains the enhanced reactivity of the encapsulated complex.
Among the different support materials, zeolites are the obvious choice for
anchoring metal complexes. Their pore diameter and pore geometry introduce size
and shape selectivity in reaction. The advantage of enzyme catalysis is that it
facilitates reactions in mild conditions. They work well in dilute aqueous solutions
and at moderate temperature, pressure and pH, and provide energy-efficient routes
for carrying out a particular reaction. The reactions catalyzed are by and large
environmentally friendly in that by-products and waste materials are a minimum.
The catalysts and the materials they synthesize are generally biodegradable. The
desire to exploit the enzyme-like property of zeozymes has prompted researchers
to synthesize complex molecules embedded in inorganic molecules so that they
can be employed in synthetically useful reactions.
1.3. Scope of the present study
Selective oxidation reactions are of prime importance in the synthesis of
fine chemicals and a variety of other processes in chemical industry. Transition
metal complexes often serve as effective catalysts in selective oxidation reactions.
But neat transition metal complexes have drawbacks, as they show loss in activity
23
and difficulty of separation, and necessitate expensive regeneration processes. It is
possible to enhance the activity, selectivity, stability, and ease of separation of a
transition metal complex and reduce the regeneration cost by encapsulating it in a
zeolite support. The zeolite supports are known to play an important role in
dispersing the active metal complex and influencing its electronic and adsorption
properties, thereby changing the catalytic properties.
The present study was undertaken with a view to
synthesizing and characterizing new zeolite encapsulated
complexes which might work as good eco-friendly catalysts;
examining the effect of the zeolite framework on the structure and
geometry of the complex;
making a comparative study of the neat complexes and zeolite
encapsulated complexes as catalysts;
assessing the catalytic activity of the synthesized complexes in
(i) the decomposition of hydrogen peroxide
(ii) the selective hydroxylation of phenol to
hydroquinone, and
(iii) the partial oxidation of benzyl alcohol to
benzaldehyde;
substituting aqueous hydrogen peroxide as oxidant in place of
organic oxidants / concentrated hydrogen peroxide, thus making use
of a clean, safe, and enviro-economic system;
examining the correlation between the activity and the structure of
the complex;
assessing the shape-selective effect of the catalysts;
studying the reaction by varying the different parameters; and
verifying the poison resistance and recyclability of the catalysts.
With these objectives, some new zeolite encapsulated complexes of
iron(III), cobalt(II), cobalt(III), nickel(II), copper(n) and palladium(II) were
24
synthesized and characterized. The ligands used for the preparation of the
complexes were:
1) 2-styrylbenzimidazole (SB)
2) morpholine-N-carbodithioate (MDTC)
3) N-ethyl-N-phenyl dithiocarbamate (EPDTC)
4) di-iso-propyl dithiocarbamate (IPDTC)
5) dimethylglyoxime (DMG)
6) salicylaldehyde semicarbazone (SALSC)
7) NN' -bis(3 -pyridylidene )-1 ,2-phenylenediamine (SPO)
8) N,N' -bis(3-pyridylidene)-1 ,4-phenylenediamine (SPP)
Some of the encapsulated complexes of copper and palladium are found to
be active catalysts for the decomposition of hydrogen peroxide, hydroxylation of
phenol to hydroquinone, and the partial oxidation of benzyl alcohol to
benzaldehyde. Further advantages of these catalysts are the easy separation of the
catalyst, possibility of recycling and amenability to continuous processing.
25
References
1. P. Baumeister, M.Studer, and F. Roessler in Handbook a/Heterogeneous
Catalysis, Vol. 5, G. Ertl, H. Knozinger, and 1. Weitkamp (Eds.), Wiley-
VCH, New York, 1997,2252.
2. RA. Sheldon, Catal. Today 1, 5 (1987) 351.
3. P.A. Sermon, Y. Sun, and KM. Keryon, Catal. Today 17 (1993) 391.
4. C.O. Wood and P.E. Garrou, Organometallics 3 (1984) 170.
5. T.O. Manly, Chem. Ind. London (1962) 12.
6. 1.P. Schinnann and S.Y. Oelavarenne, Hydrogen Peroxide in Organic
Chemistry, Information Chimie, Paris, 1979.
7. 1.H. Baxendale, Adv. Cata!' 4 (1952) 31.
8. E. Tsuchida and H. Nishide, Adv. Polym. Sci. 24 (1977) 1.
9. R Sreekala and KK Mohammed Yusuff, Ind. J. Chem., 34A (1995) 994.
10. H. Sigel,Angew. Chem. 81 (1969) 161.
11. H. Sigel, Angew. Chem., Int. Ed. Engl. 8 (1969) 167.
12. P. Jones and I. Wilson, Met. Ions BioI. Syst. 7 (1978) 185.
13. 1. Schubert, VS. Shanna, E.R. White, and L.S. Bergelson, J. Am. Chem.
Soc. 90 (1968) 4476.
14. 1. Schubert and VS. Shanna, Proc. Int. Con! Coord. Chem., 1 t
h
(1968) 19.
15. V.S. Sharma and 1. Schubert,lnorg. Chem., 10,2 (1971) 251.
16. Chandra R. Jacob, Saji P. Varkey, and Paul Ratnaswamy, Appl. Cata!' A
168,2 (1998) 353.
17. Jayendra Patole, Sabari Outta, Subhash Padhye, and Ekkehard Sinn, Inorg.
Chim. Acta 318 (2001) 207.
18. M. Krajnem and 1. Lever,Appl. Cala!. B. 3,2-3 (1994) 101.
19. C.E.K Mees, The Theory a/the Photographic Processes, Vol. 33,
Macmillan, New York, 1942.
20. F. Bourdin, M. Costantini, M. Jouffret, and G. Latignan, Ger. Patent 2 064
497, 1971.
21. G.A. Hamilton, 1. P. Friedman, and P.M. CampbeU, J. Am. Chem. Soc. 88
(1966) 5266.
22. K Hotta, S. Tamagaki, Y. Suzuki, and W. Tagaki, Chem. Lett. (1981) 789.
23. S. Tamagaki, M. Sasaki, and W. Tagaki, Bull. Chem. Soc. Jpn. 62 (1989)
153.
24. I. Yu Litivinsev, Yu V Mitnik, A. I. Mikhailyuk, S. V Timofeev, and N. V
Sapunov, Kinet. Kala!. (1) 34 (1993) 56.
25. P. Maggioni, US Patent 391 432, 1975.
26. P. Maggioni and F. Minisci, Chim. Ind. 59 (1977) 239.
27. P. Maggioni, US Patent 3 929913,1975.
28. D.RC. Huybrechts, I. Vans en, H.x. Li, and P.A. Jacobs, Catal. Lett. 8
(1991) 237.
26
29. S. Lin, Y Zhen, S.M. Wang, and YM. Dai, J. Mol. Cata!' 156 (2000) 113.
30. U. Romano, A Esposito, F. Maspero, e. Neri, and M.G. Clerici, Stud. SlIlj
Sci. Catal. 55 (1990) 33.
31. e.D. Chang and S.D. Hellring, US Patent 4578521, 1986.
32. Felipe Alagarra, M Alexandra Esteves, Vincente Fomes, Herminegildo
Garcia, and Jalme Primo, New J. Chem. (1998) 333.
33. Rolfe Fricke, Hendrik Kesslick, Giinter Lischke, and Manfred Richter,
Chem. Rev. 100 (2000) 2303.
34. H.S. Bloch, US Patent 3 580 956, 1971.
35. A. Tuel, S.M. Khouzami, YB. Taarit, and e. Naccache, J. Mol. Catal. 68
(1991) 45.
36. lA. Martens, P.H Buskens, P.A. Jacobs, A van der Po I, J.He. van Hooff,
e. Ferrini, HW. Kouwenhoven, PJ. Kooyman, and H van Bekkum, Appl.
Catal. A 99 (1993) 71.
37. A. Esposito, M.Taramasso, and e. Neri, US Patent 4396783, 1983.
38. A. Esposito, M.Taramasso, C. Neri, and F. Buonomo, British Patent 2
116974,1985.
39. lE. Backvall, AK Aswathi, and Z.D. Renko, J. Am. Chem. Soc. 109
(1987) 4750.
40. lE. Backvall, Stud. Surf. Sci. Cata!. 41 (1988) 105.
41. lE. Backvall, RB. Hopkins, H. Grennberg, M. Mader, and AK Aswathi,
J. Am. Chem. Soc. 112 (1990) 5160.
42. A.Puzari, L. Nath, and Jubaraj B. Baruah, Ind. J. Chem. 37 A (1998) 723.
43. R Sumathi, K Johnson, B. Viswanathan, and T.K Varadarajan, Ind. J.
Chem. 38 A (1999) 40.
44. Won K Seok, Bull. Korean Chem. Soc. 20,4 (1999) 395.
45. A. Corma and H Garcia, Chem Rev. 102 (2000) 3837.
46. le. Bailar, Jr., Cat. Rev. Sci. Eng. 10 (1974) 17.
47. T. Joseph, D.P. Sawant, e.S. Gopinatb, and S.B. Halligudi, J. Mol.
Cata1.184, 1-2 (2002) 289.
48. A. Kozlov, A Kozlova, K Asakura, and Y Iwasawa,. J. Mol. Cata!' 137, 1-
3 (1999) 223.
49. lM. Sabater, A Corma, A Domenech, V. Fomes, and H Garcia, Chem.
Commun. 14 (1997) 1285.
50. B. Fan, R Li, W. Fan, l Cao, and Z. Bing, J. Natural Gas Chemistry 9,2
(2000) 157.
51. S.A. Chavan, D. Srinivas, and P. Ratnasamy, Chem. Commun. (Cambridge,
United Kingdom) 12 (2001) 1124.
52. S.N. Rao, KN. Munshi, and N.N. Rao,J. Mol. Cata!' 145, 1-2 (1999) 203.
53. RF. Parton, Ivo F.J. Vankelecom, M.J.A Casselman, e.P. Bezoukhanova,
lB. Uytterhoeven, and P.A Jacobs, Nature 370 (1994) 541.
54. Agnes Zsigmond, Ferenc Notheisz, and lE. Backvall , Catal. Lett. 65, 1-3
(2000) 135.
55. T. Kimura, A Fukuoka, and M. Ichikawa, Catal. Lett. 4, 4-6 (1990) 279.
56. S. Seelan, D. Srinivas, M.S. Agashe, N.E. Jacob, and S. Sivasankar, Stud.
SUIt Sci. Catal. 135 (2001) 2224.
27
57. R. Raja and P. Ratnasamy, Stud. Suif. Sci. Catal. 105 (1997) 1037.
58. M. Ichikawa, T. Kimura, and A FukUoka, Stud. Surf Sci. Catal. 60 (1991)
335.
59. N. Herron, J. Coord. Chem. 19 (1988) 25.
60. R. Raja and P. Ratnasamy, Stud. Suif. Sci. Catal. 101 (I996) 181.
61. R. Raja, Chandra R Jacob, and Paul Ratnasamy, Catal. Today 49, 1-3
1999)171.
62. Chris Bowers and Prabir K. Dutta, J. Catal. 122,2 (1990) 271.
63. Saji P. Varkey, Chandra Ratnasamy, and Paul Ratnasamy, J. Mol. Catal.
135,3 (1998) 295.
64. Saji P. Varkey and Chandra R Jacob,Ind. J. Chem. 37A, 5 (1998) 407.
65. LW.CE. Arends, Pellizon M. Birelli, and RA Sheldon, Stud. Surf. Sci.
Catal. 11 0 (1997) 1031.
66. R. Raja and P. Ratnasamy, Catal. Lett. 48 (1997) 1.
67. K.1. Balkus, Jr., AK. Khanmamedova, M.K. Dixon, and F. Bedioui, Appl.
CatalA 143, 1 (1996) 159.
68. Trissa Joseph, CS. Sajanikumari, S. Deshpande, and S. Gopinathan,Ind. J.
Chem. 38A, 8 (1999) 792.
69. M.R Maurya, SJJ. Titinchi, S. Chand, and LM. Mishra, J. Mol. Cata1.180,
1-2 (2002) 201.
70. Saji P. Varkey and Chandra R Jacob, Ind. J. Chem. 38A, 4 (1999) 320.
71. P.P. Knops-Gerrits, D.E De Vos, F. Thibault-Starzyk, and P.A Jacobs,
Nature 369 (9194) 543.
72. P.P. Knops-Gerrits, F. Thibault-Starzyk and P.A Jacobs, Stud. Suif. Sci.
Catal. 84 (1994) 1411.
73. A.W. van de Made, J.W.H. Smeets, RJ.M. Nolte, and W. Drenth, J. Chem.
Soc., Chem. Commun. (1983) 1204.
74. R. Grommen, P.Manikandan, Y Gao, T. Shane, J.1. Shane, RA
Schoonheydt, B.M. Weckhuysen, and D. Goldfarb, J. Am. Chem. Soc. 122
(2000) 11488.
75. T. Tatsumi, M. Nakamura, and H. Tominaga, Catal. Today 6 (1989) 163.
76. S.B. Ogunwumi and T. Bein, Chem. Commun. (1997) 901.
77. r. Joseph, S.B. Halligudi, C Satyanarayanan, D.P. Sawant, and S.
Gopinathan,J. Mol. Catal. 168, 1-2 (2001) 87.
78. A.F. Cronstedt, Akad. Handl. 17 (1756) 120.
79. M.A. Damour, Ann. Chimie. 3, 53 (1858) 438.
80. LG. Kovzun, YL Tarasevich, YY. Maslyakevich, and AI. Zhukova, L7cr.
Khim. Zh. 43 (1977) 247.
81. John Meurig Thomas, Osamu Terasaki, L. Pratibha Gai, Wuzong Zhou, and
Jose Gonzalez - Calbet, Acc. Chem. Res. 34 (2001) 583.
82. S.T. Wilson, B.M. Lok, CA Messina, I.R Cannan, and E.M. Flanigen, J.
Am. Chem. Soc. 104 (1982) 1145.
83. M.W. Anderson, O. Terasaki, T. Ohsuna, I. Philippou, S.P. Mackay, J.
Rocha, and S. Lidin, Nature 367 (1994) 347.
84. 1. Dwyerand K. Karim,J. Chem. Soc., Chem. Commztn. (1991) 905.
85. Cr. Kresge, M.E. Leonowicz, WJ. Roth, J.C Vartuli, and J.S. Beck,
28
Nature 359 (1992) 710.
86. lS. Beck, J.C. Vartuli, WJ. Roth, M.E. Leonowicz, c.T. Kresge, K.D.
Schmidt, c.T.W. Chu, D.H. Olson, E.W. Sheppard, S.B. Mc Cull en, lB.
Higgins, and J.L. Schlenker, J. Am. Chem. Soc. 114 (1992) 10834.
87. R.M. Barrer, Hydrothermal Chemistry ojZe 0 lites , Academic Press, New
York,1982.
88. R.M. Barrer, Hydrothermal Chemistry ojZeolites, Academic Press, London,
1983.
89. WJ. Mortier, Compilation oj Extra Frame'work Sites in Zeolites,
Butterworths, London, 1982.
90. R.M. Barrer, Zeolites and Clay Minerals as Sorbents and Molecular Sieves,
Academic Press, London, 1978.
91. lV Smith, Adv. Chem. Ser. 101 (1971) 17l.
92. lV Smith, Chem. Rev. 88 (1988) 149.
93. lA Rabo, Zeolite Chemistry and Catalysis, A CS Monograph 171,
American Chemical Society, Washington DC, 1976.
94. Young Hoon Yeom, Yang Kim, Seung Hwan Song, and Karl Seff,J. Phys.
Chem. 101 (1997) 2140.
95. P.G. Menon, in Lectures on Catalysis, 41 si Annual Meeting, Ind. Acad.
Sci., S. Ramasheshan (Ed.) 43, 1975.
96. A Corma and Augstin Martinez, Adv. Mater. 7,2 (1995) 137.
97. N. Herron, Inorg. Chem. 25 (1986) 4714.
98. C. Bowers and P.K. Dutta, J. Catal. 112 (1990) 271.
99. L. Gail1ion, N. Sajot, F. Bedioui, 1. Devynck, and K.1. Balkus, Jr., J.
Electroanal. Chem. 345 (1993) 157.
100. K.l Balkus, Jr., AA Welch, and B.E. Gnade, Zeolites 19 (1990) 722.
101. S. Kowalak, RC. Weiss, and K.J. Balkus, Jr., J. Chem. Soc., Chem.
Commun. (1991) 57.
102. F. Bedioui, L. Roue, 1. Devynck, and K.1. Balkus, Jr., Stud. Suif. Sci. Cata!'
84 (1994) 917.
103. D.E. De Vos, F. Thibault-Starzyk, and P.A Jacobs,Angew. Chem. 1nt. Ed.
Engl. 33 (1994) 432.
104. W.H. Quale, G. Peeters, G.L. De Roy, E.F. Vansant, and J.H. Lunsford,
Inorg. Chem. 21 (1982) 2226.
105. W. De Wilde, G. Peeters, and H. Lunsford, J. Phys. Chem. 84 (1980) 2306.
106. P.P. Knops-Gerrits, D.E. De Vos, F. Thibault-Starzyk, and P.A Jacobs,
Nature 369 (1994) 543.
107. B.V Romanovsky, in Proc. 8
th
Int. Congr. Ca tal. , Vol. 4, Verlag Chemie,
Weinheim 1984,657.
108. VYZakharov and B.V Romanovsky, Vest. Mosk. Univ., Ser. 2 Khim., 18
(1977) 142.
109. B.V. Romanovsky and AG. Gabrielov, J. Mol. Catal. 74 (1992) 293.
110. G. Meyer, D. Wohrle, M. Mohl, and G. Schulz-Ekloff, Zeolites 4 (1984)
30.
Ill. KJ. Balkus, Jr., and lP. Ferraris, J. Phys. Chem. 94 (1990) 8019.
112. E. Paez-Mozo, N. Gabriunas, F. Lucaccioni, D.D. Acosta, P. Patrono, A La
29
Ginestra, P. Ruiz, and B. Delmon, J. Phys. Chem. 97 (1993) 12819.
113. lP. Ferraris, KJ. Balkus, Jr., and A Schade, J. Incl. Phenom. 14 (1992)
163.
114. N. Herron, G.D. Stucky, and C.A Tolman, J. Chem. Soc., Chem. Commlln.
(1986) 1521.
115. T. Kimura, A Fukuoka, and M. Ichikawa, Catal. Left. 4 (1990) 279.
116. Ziqi Jiang and Zuwei Xi, Fenzi Cuihua 6,6 (1992) 467.
117. M. Ichikawa, T. Kimura, and A Fukuoka, Stud. Surf. Sci. Cata!' 60 (1991)
335.
118. AG. Gabrielov, K1. Balkus, Jr., S.L. Bell, F. Bedioui, and 1. Devynck,
Microporous Materials 2 (1994) 119.
119. F. Bedioui, L. Roue, L. Gaillion, 1. Devynck, S.L. Bell, and K1. Balkus, Jr.,
Preprints Div. of Petroleum Chemistry, ACS 38,3 (1993) 529.
120. KJ. Balkus, Jr., AG. Gabrielov, S.L. Bell, F. Bedioui, L. Roue, and 1.
Devynck,Inorg. Chem. 33 (1994) 67.
121. M. Nakamura, T. Tatsumi, and H. Tominaga, Bull. Chem. Soc. Jpn. 63
(1990) 3334.
122. YW. Chan and RB. Wilson, Preprints Div. of Petroleum Chemistry, ACS
33,3 (1988) 453.
123. l Woltinger, lE. Backvall, and A Zsigmond, Eur. Chem. J. 5 (1999) 2084.
124. Agnes Zsigmond, Krisztilin Bogar, and Ferenc Notheisz, J. Cata!' 213
(2003) 103.
125. L.A Rankel and E.W. Valyocsik, US Patent, 4500503, 1985.
126. G. Schulz-Ekloff and S. Emst in Handbook of Heterogeneous Catalysis
Vol. 1, G. Ertl, H. Knozinger, and 1. Weitkamp (Eds.), Wiley-VCH, New
York, 1997,374.
127. Chen-Ho Tung, Li-Zhu Wu, Li-Ping Zhang, and Bin Chen, Acc. Chem. Res.
36 (2003) 39.
128. l Haber, Pure and Appl. Chem. 66, 8 (1994) 1597.
129. Isao Mochida, Akio Kato, and Tetsuro Seiyama, Bull. Chem. Soc. Jpn. 45
(1972) 2230.
130. A Maldotti, A Molinari, and R Amadelli, Chem. Rev. 102 (2002) 3811.
131. X.Z. Qin and AD. Trifunac, J. Phys. Chem. 94 (1990) 4751.
132. K.B. Yoon, Chem. Rev. 93 (1993) 321.
133. B.v. Romanovsky, VY. Zakharov, and T.G. Borisava, Moscow Univ. Pub!.
(1982) 170.
134. lH. Lunsford, Catal. Rev. -Sci. Eng. 12 (1975) 137.
135. S. Chavan, D. Srinivas, and P. Ratnasamy, J. Catal. 192,2 (2000) 286.
136. K.O. Xavier, 1. Chacko, and KK Mohammed Yusuff, J. Mol. Cata!' 178,
(2002) 275.
137. G.A Ozin and Caroline Gil, Chem. Rev. 89 (1989) 1749.
138. Chandra R Jacob, Saji P. Varkey, and Paul Ratnasamy,Appl. Cata!' A 168
(1998)353.
139. S. Deshpande, D. Srinivas, and P. Ratnasamy,J. Catal. 188,2 (1999) 26l.
140. A Domenech, P. Formentin, H. Garcia, and M.J. Sabater, European J.
Inorg. Chem. 6 (2000) 1339.
30
141. N. Herron, G.D. Stucky, and C.A Tolman, Inorg. Chim. Acta. 100 (1985)
135.
142. N. Herron, New J. Chem. 13 (1989) 761.
143. N. Herron, Chem. Tech. Sept. (1995) 542.
144. D.E. De Vos, F. Thibault-Starzyk, P.P. Knops-Gerrits, RF. Parton, and P.A
Jacobs,Macromol. Symp. 80 (1994) 157.
145. IT. Groves and T.E. Nemo, J. Am. Chem. Soc. 105 (1983) 5786.
146. N. Herron and C.A Tolman, J. Am. Chem. Soc. 109 (1987) 2837.
147. RF. Parton, L. Uytterhoeven, and P.A Jacobs, Stud. Surf. Sci. Catal. 59
(1991) 395.
148. lvo F.J. Vankelecom, RF. Parton, J.AC. Mark, J.B. Uytterhoeven, and P.A
Jacobs, J. Catal. 163 (1996) 457.
149. lvo F.J. Vankelecom, Chem. Rev. 102 (2002) 3779.
150. F.M.P.R van Laar, F. Holsteyns, I.F.J. Vankelecom, S. Smeets, W. Dehaen,
and P.A Jacobs, J. Photochem. Photobiol. A. Chem. 144 (2001) 141.
15l. RF. Parton, I.F.J. Vankelecom, D. Tas, K. Janssen, P.P. Knops-Gerrits, and
P.A Jacobs,J. Mol. Catal. 113, 1-2 (1996) 283.
152. S. Chavan, D. Srinivas, and Paul Ratnasamy, Topics in Catalysis 11/12, 1-4
(2000) 359.
31
Chapter 2
Materials and methods
2.1. Introduction
This chapter presents details of reagents and other materials used in the present
study. It describes the methods used for the synthesis of ligands, simple complex,
metal exchanged zeolite Y, and zeolite Y encapsulated complexes. The various
physico-chemical methods employed for the characterisation of the zeolite Y
encapsulated complexes are discussed. Details of the different instruments and
methods used for the study of the catalytic activity of the compounds prepared are
also given.
2.2. Reagents
The following metal salts were used: Fe(N0
3
)3 9H
2
0 (S.D. Fine. Chem.
Ltd. GR); CoCh 6H
2
0 (E. Merck, GR); NiCh 6H
2
0 (E. Merck, GR); CuCh
2H
2
0 (E. Merck, GR); PdCh ( SRL GR); NaCI (E. Merck , GR).
Zeolite Y with a Si/AI ratio 2.4 and particle density 2.5 x 10
3
kg/m3,
obtained from Sud-Chemie India Ltd., Binanipuram, Ko chi, was used as support.
Dimethylglyoxime (E. Merck); salicylaldehyde (E. Merck); ethylene glycol
(Aldrich); semicarbazide hydrochloride (Loba Chemie Pvt. Ltd.); pyridine-3-
carboxaldehyde (E. Merck); a-phenylenediamine (BDH); p-phenylenediamine
(CDH); cinnamic acid (Aldrich); carbon disulphide (CDH); N-ethylaniline (E.
Merck); morpholine (SRL); di-iso-propylamine (Riedel-De Haen AG Seelz-
Hannover) were used for the syntheses of ligands.
Hydrogen peroxide (30% w/v, E. Merck), benzyl alcohol (SRL) and phenol
(Nice Chemicals Pvt. Ltd.) were used for catalytic studies on simple and supported
complexes. Unless otherwise specified, all reagents used were of analytical reagent
grade. The solvents used were either of 99% or purified by known laboratory
procedures.
33
Liquid nitrogen (Sterling gases, Ko chi) was used for surface area
measurements. Air, nitrogen, and hydrogen gas cylinders (Sterling gases, Kochi)
were used for catalytic studies.
2.3. Synthesis of Iigands
2.3.1. Preparation of dimethylglyoxime (DMG)
Dimethylglyoxime was purified by recrystallising it from methanol. The pure
crystals were dried over anhydrous calcium chloride.
2.3.2. Synthesis of 2-styrylbenzimidazole (SB)
Styrylbenzirnidazole was prepared following the procedure suggested by
Dubey and Ramesh [1]. The synthesis was carried out in two stages. Details of the
preparation are given below.
a) Preparation of o-phenylenediamine dihydrochloride
A solution containing concentrated hydrochloric acid (60 rnI), water (40 rnI)
and SnCh (2 g) was prepared, and a-phenelenediarnine was added to it. To this
solution decolourising carbon was added and it was heated to boiling. The solution
was filtered, and concentrated hydrochloric acid was added to it. Finally it was
cooled in freezing mixture. The crystals that separated were filtered, washed with
hydrochloric acid, and dried in a vacuum desicator over sodium hydroxide [2].
b) Synthesis of styrylbenzimidazole
A mixture of a-phenylenediamine dihydrochloride (1.81g, lOmrnoles) and
cinnamic acid (1.48 g, 10 mrnoles) in ethylene glycol (lOml) was refluxed for 5
34
hours. The reaction mixture was cooled to room temperature and poured into water
(100ml). The solid separated was filtered, washed with water, dried, recrystallised
using methanol, and fmally dried over anhydrous CaCh [I].
2.3.3. Synthesis of carbodithioate and dithiocarbamates
Morpholine-N-carbodithioate (MOTC), N-ethyl-N-phenyldithiocarbamate
(EPDTC) and di-iso-propyldithiocarbamate (IPDTC) were synthesized using the
following procedure [3]. Sodium hydroxide (20g, 0.5moles) and the amine
(morpholine, di-iso-propylamine, or N-ethylaniline) (0.5mole) were mixed together
in a 500ml two-necked flask. This was cooled in a freezing mixture of ice and salt.
Carbon disulphide (3Iml, 0.5moles) was added drop-wise from a separating funnel
stirring the mixture using an electric stirrer. The addition was carried out very
slowly, in nearly two hours. The solid that separated was washed several times with
hexane and recrystallised from water and dried over anhydrous calcium chloride.
2.3.4. Synthesis of N,N'-bis(3-pyridyJidene)-1,2-pheneJenediamine (SPO)
The Schiff base N,N'-bis(3-pyridylidene)-1,2-phenelenediamine was
prepared by refluxing pyridine-3-carboxaldehyde (2.14g, 0.02mole) and 0-
phenylenediamine (1.08g, O.Olmole) for five hours. The yellow product was filtered
and washed with benzene several times. It was recrystallised from absolute alcohol
and dried over anhydrous calcium chloride [4].
2.3.5. Synthesis of N,N'-bis(3-pyridyJidene)-1,4-phenyJenediamine (SPP)
The ligand was prepared in the same way as N,N'-bis(3-pyridylidene)-1,2-
phenelenediamine, using p-phenylenediamine in the place of a-phenylenediamine,
the quantities of reagents being the same. The product was filtered, washed with
35
benzene, and recrystallised from alcohol. It was dried over anhydrous calcium
chloride.
2.3.6. Synthesis of salicylaldehyde semicarbazone (SALSC)
This ligand was prepared using the procedure suggested by Jayendra Patole
et al. To an aqueous solution of semicarbazide hydrochloride (lg, O.OOSmole) and
sodium acetate (l.Sg, O.Olmole), salicylaldehyde (O.Sg, 0.004mole) was added with
stirring. The mixture was refluxed on a water bath for 2 hours. The semicarbazone
separated was ftltered, washed with ethanol, and recrystallised from I-propanol [S].
2.4. Synthesis of zeolite Y encapsulated metal complexes
2.4.1. Preparation of sodium exchanged zeolite (NaY)
The standard ion exchange procedure was used for the preparation of NaY.
Zeolite Y (HY, Sg) was mixed with NaCI solution (0.1, SOOml) and stirred for 24
hours at room temperature. This was done to remove extra framework iron and ion
exchangeable impurities. The solution was ftltered and washed with deionised water
till the filtrate was free of chloride ions. The NaY formed was dried at 100C for 2
hours. The procedure was based on the one reported by Edward et al. [6] with
slight modification.
2.4.2. Preparation of transition metal ion exchanged zeolite (MY)
(M = Fe, Co, Ni, Cu, Pd)
Sodium exchanged zeolite (S.Og) was stirred with metal salt solutions of
ferric nitrate and chlorides of cobalt, nickel, copper, and palladium (SOO ml; O.OIM
I O.OOIM) at 90C for 12 hours. Low concentration of metal salt solutions of pH
4.0-4.5 was used as dealumination occurs at higher concentrations [7]. The slurry
36
was filtered and washed with deionised water to make it free from anions. It was
dried at 100 QC for 2 hours and finally dehydrated at 450 QC for 4 hours [8].
2.4.3. Encapsulation of metal complexes in zeolite
The flexible ligand method was used for encapsulating metal complexes in
the cages of the zeolite [9]. Many of the free ligand molecules are flexible enough
to pass through the restricting windows giving acces; to the super cages. These
ligands can react with the metal ions already present in the super cages, and
encapsulated complexes can be formed. Two general procedures are followed for
this kind of encapsulation.
2.4.3.1. Encapsulation by heating in a sealed ampule
Metal exchanged zeolite MY (3. Og) was mixed thoroughly with excess of
ligand (ligand to metal mole ratio -2 - 4). It was placed in a glass vial and the open
end of the tube was sealed so as to form an ampule. It was then heated in a furnace
at an optimum temperature and for a definite period. This allows the sorbate
molecules to disperse uniformly throughout the sample so that complexation occurs
effectively. Complex formation involves physical entrapment of metal complexes in
the cages of the zeolite rather than covalent or ionic attachment to the walls.
Because of size restriction the complex stays in the zeolite and cannot diffuse out,
provided the complex is stable. The product obtained was soxhlet extracted using a
suitable solvent to remove excess ligand and surface species in cases where the
complex and the ligand are soluble in that particular solvent. When the solubilities
of the complex and ligand are different, they have to be soxhlet extracted with the
suitable solvents. The soxhlet extraction was continued until the extract becomes
colourless indicating complete removal of the species to be eliminated. The
uncomplexed metal remaining in the zeolite was removed by back exchange of the
37
zeolite with NaCI solution (O.OIM, 250ml) for 24 hours. It was then filtered,
washed free of chloride ions, and fmally dried at 100C for 2 hours [10].
2.4.3.2. Encapsulation by refluxing the metal exchanged zeolite with the ligand
Metal exchanged zeolite MY (3.0g) was added to a solution of the ligand
(excess ligand should be present) in a suitable solvent (ligand to metal mole
ratio-2-4). The mixture was refluxed for 10 hours over a water bath. The ligand
penetrates through the charmels of zeolite and complexes are formed. Soxhlet
extraction, back ion exchange, and removal of chloride are to be done in this case
also for the removal of excess ligands and surface complexes.
2.5. Synthesis of neat complex
Cupric chloride was mixed with styrylbenzimidazole in methanol. It was
refluxed for 3 hours. The crystals that separated were flltered, and dried using
anhydrous calcium chloride.
2.6. Catalysis procedures
2.6.1. Decomposition of hydrogen peroxide
The catalytic activity of the synthesized samples was evaluated by
monitoring the volume of oxygen evolved from the liquid phase using a gas burette
attached to the reaction flask. The rate of oxygen gas evolution was monitored at
room temperature and atmospheric pressure using gas burette. The volume of the
evolved gas was monitored at intervals of 5 minutes in all the cases. The procedure
is given in detail in Chapter 7.
38
2.6.2. Hydroxylation of phenol to hydroquinone
The hydroxylation of phenol was carried out by stirring a mixture of phenol,
catalyst, water and hydrogen peroxide. The reaction mixture was filtered to remove
the catalyst, and the product obtained was analysed using gas chromatograph. The
quantities of the reactants used and detailed procedure for each reaction are given
in Chapter 8.
2.6.3. Partial oxidation of benzyl alcohol to benzaldehyde
A mixture of benzyl alcohol, hydrogen peroxide, the activated catalyst,
and solvent (water) was stirred in a reaction flask of SOml capacity for a definite
period of time using a magnetic paddle. The aqueous and organic layers were
separated. The organic layer was analysed using thin layer chromatography (TLC)
to detect the formation of benzaldehyde. It was finally analysed using gas
chromatography to quantify the amount of substrate and product. The detailed
procedure and the quantities of the reactants used for each reaction are given in
Chapter 9.
2.7. Characterization methods
A variety of techniques have been developed for the characterisation of the
catalysts and to trace the course of catalytic reactions. Characterization of the
zeolite sample is carried out to get a better understanding of encapsulation, the
physico-chemical properties of the complex, distribution of the complex in the host,
the influence of the guest host interaction on the structure of the complex, the
stability of the encapsulated complex compared to that of the neat complex,
retaining of the zeolite framework even after encapsulation, and so on. The
following techniques are generally used for the characterization of neat and zeolite
encapsulated samples. The various techniques are:
39
i) chemical analysis
ii) surface area and pore volume analysis
iii) x-ray diffraction studies
iv) spectroscopic methods
v) magnetic susceptibility measurements
vi) scanning electron microscopy
vii) thermal analysis
viii) conductometric measurement
2.7.1. Chemical analysis
2.7.1.1. Analysis of Si, AI, Na, and transition metal ion in the zeolite sample
Analyses of Si, AI, Na and the transition metal ions are done by destroying
the zeolite lattice-using con. H
Z
S0
4
A known weight (Wl) of the sample was
taken in a beaker, and treated with concentrated sulphuric acid (95%, 40 ml) and
was heated until S03 fumes were evolved. It was then cooled, diluted with Vo-ater
and filtered using ash-less filter paper. The filtrate was collected in a standard
flask. The residue was heated in a platinum crucible and was weighed again (wz).
To the calcinated material 40% HF was added in drops, warmed and strongly
heated to dryness. The procedure was repeated five to six times. During this
process Si is converted into H
z
SiF
6
. It was again incinerated to 800 QC for 1 hour,
cooled and weighed (W3). From the loss in weight the amount of silica present in
the sample can be estimated using the equation:
% SiO
z
= (W3-WZ) x 100/ Wl
The residue was then fused with potassium persulphate (2-3 g) to form a clear
molten mass. The mass when cooled was dissolved in water, and this was added to
the filtrate collected in the standard flask. This solution was analysed for AI, Na and
metal contents usmg Perkin Elmer Model 3110 Atomic Absorption
40
Spectrophotometer. Knowing Si, Al, Na, and transition metal mole ratio it IS
possible to get the unit cell formula of ion exchanged zeolite.
2.7.1.2. Determination of metal in neat complex
A known weight (0.20-0.30g) of the complex was treated with concentrated
sulphuric acid (Srnl) followed by concentrated nitric acid (20ml). After the reaction
subsided perchloric acid (60%; Sml) was added. This solution was maintained at
boiling temperature for 3 hours on a sand bath. The clear solution thus obtained
was evaporated to dryness, cooled, treated with concentrated nitric acid (Sml) and
was again evaporated to dryness on a water bath. The residue was dissolved in
water and was used for the estimation of metals. Iodometric method was employed
for the estimation of copper in the complex [11].
2.7.1.3. Determination of chlorine
Chloride estimation was carried out as follows [11]. In a nickel crucible
a layer of Na
2
C0
3
was taken, and a layer of Na
2
02 was set above the layer of
Na2C03. A definite quantity of the sample, accurately weighed out, was placed over
this. This was then covered with another layer of Na
2
02 and then with a layer of
Na
2
C0
3
. The mixture was heated to get a fused mass. The nickel crucible with
substance was immersed in water taken in a beaker and heated over a water bath
until the substance dissolved completely, and then the crucible was taken out. The
solution was neutralised with 1: 1 HN0
3
, adding the acid till effervescence ceased
and then in slight excess. Any residue that was present was filtered off and the
chloride in the solution was estimated using Volhard method [11].
41
2.7.2. CHN analysis
Microanalyses of carbon, hydrogen and nitrogen were done at the
Sophisticated Instrumentation Centre for Applied Research and Testing, Anand,
Gujarat. The CRN analyses were done on a Carto Erba Analyser Model 1108. The
results were related to the amounts of metal obtained from AAS for the
encapsulated complexes. This helped to fmd the amount of uncomplexed metal ions
if any, which remain in lattice.
2.7.3. Surface area and pore volume analysis
It is necessary to know the surface area of a solid for catalytic application.
This fundamental property of the solid is called surface area and is generally given
as m
2
/g of solid. The surface areas of the zeolite samples were determined by the
BET method [12] of nitrogen adsorption at liquid nitrogen temperature using
'Micromeritics Gemini 2360'. The weighed amount of sample is taken in a dry,
clean sample tube. The volume of gas adsorbed by the sample was monitored at
different relative pressures in the range 0.1-0.9. The zeolite sample was heated to
473 K and maintained at this temperature for 3 hours in a stream of dry nitrogen
atmosphere (50 ml/h) to remove any volatile components from the surface of the
solid. After this pre-treatment the sample was cooled to room temperature. The
sample tube together with substance was weighed and was fIxed to the instrument
with the help of sample holders. It was then brought to 77 K using liquid nitrogen
as coolant and mixture of nitrogen and hydrogen was passed over the catalyst.
Only nitrogen was physically adsorbed on the surface of the solid at this liquid
nitrogen temperature. This decreases the pressure in the chamber until the adsorbed
gas is in equilibrium with the free gas phase.
Surface area is calculated using the following BET equation
where
42
=
C
=
volume of gas adsorbed at relative pressure PlP
o
saturated vapour pressure
volume of gas adsorbed for monolayer coverage
BET constant
By plotting the left side of the BET equation against PIP 0 (up to 0.3), a straight line
is obtained with a slope of(C-1)N
m
C and an intercept 1N
m
C. From these values,
V m and hence X
m
, the number of moles of N2 adsorbed, can be calculated.
BET surface area is calculated using the equation
S BET = Xm N Am 1 0-
20
where
N = Avogadro's number
Am = cross-sectional area of the adsorbate molecule in A2
The total pore volume of the sample at PlPo - 0.9 is computed by
converting the volume ofN
2
adsorbed at PlPo - 0.9 to the volume ofliquid
equivalent to it using the following equation:
VIOl = Vads D where
VIOl = total pore volume at PlPo - 0.9
Vads = volume of gas adsorbed at relative pressure -0.9
D = density conversion factor
2.7.4. X-ray diffraction studies
X-ray diffraction can be used to determine the zeolite's internal structure.
The method serves as a basic fmger printing technique, as two substances cannot
have absolutely identical diffraction patterns. The patterns of the sample in the
present study were obtained on a 'Philips P W 1710 / Rigaka D-Max C' X-ray
diffractometer. The measurements were carried out with X-ray source, Ni filtered
CuKa. radiation with A = 1.5404 and a movable detector which scans the intensity
of the diffracted beam as a function of the angle 28 between the incident and the
43
diffracted beams. The range of 28 used was between 5 and 70. The XRD patterns
of the ion exchanged zeolite and zeolite encapsulated complex are compared with
that of parent zeolite (NaY) to know whether there is any loss in crystallinity.
2.7.5. Spectroscopic analysis
Spectroscopic methods are employed as a reliable means for the
characterization of complexes and the study of catalysis. They provide information
regarding the nature and structure of the complex, and the reactivity of the surface.
Further the methods can be easily adopted for insitu studies to gain vital
information on reaction mechanism.
2.7.5.1. Electronic spectra
The characterisation technique is based on the electronic transition in the
UV-VIS-NlR region of the electronic spectrum. Usually diffuse reflectance
spectroscopy is made use of for the analysis of zeolite samples. It gives
information regarding d-d transitions, intra ligand bands, charge transfer spectra,
structure of complexes, kinetics of reactions etc.
The diffuse reflectance spectra were recorded at room temperature in the
range 200-2000nm on a Cary Win spectrophotometer at Regional Sophisticated
Instrumentation Centre, Indian Institute of Technology, Chennai and between
200run-800nm on a Perkin Elmer 320 spectrometer. NaY was used as blank for
zeolite samples and finely ground BaS04 / MgO was used as reference in the case
of neat complexes. The spectra were computer processed and plotted as
percentage reflectance versus wavelength using the principle of Kubelka-Munk
analysis [13,14,15] or as percentage absorbance versus wavelength
44
2.7.5.2. Infrared spectra
Infrared spectroscopy was used for the identification of functional groups,
and for detecting the coordinaton of ligands to transition metals. Infrared spectra of
the ligand, simple complex, metal exchanged zeolite and supported complexes in
the region 4000cm-
1
- 400cm-
1
were recorded using Schimadzu 8000 Fourier
Transform Infrared Spectrophotometer 1 Brukker IFS 66v FTIR Spectrometer. The
characteristic bands due to the coordination sites of ligands exhibit well-defined
shift upon complexation 1 chelation. But many of these bands may be masked by
strong bands of zeolite in zeolite encapsulated complexes. The framework
vibrations of zeolites are observed in the spectral range between 1250cm-
1
and
400cm-
1
. The samples were mixed with KBr, made into pellets and analysed.
2.7.5.3. EPR spectra
EPR spectroscopy has been used as a powerful tool for investigating
various physico-chernical problems such as identification of paramagnetic species,
its geometry, electronic structure, identification of catalytically active species in a
reaction and to investigate the influence of the host on the geometry of the
complex. The X-band EPR spectra of powdered samples of zeolite encapsulated
Cu(n) complexes and glass spectra of simple complexes were recorded at room
temperature and at liquid nitrogen temperature using a Varian E-109 XlQ band
spectrophotometer. The g values were estimated relative to tetracyanoethylene
( TCNE, g = 2.0027 ). The magnetic moment was determined from the EPR data
using the following equation [16]
1l
2
eff = g/14 + g/14 + 3kT/Ao ( g - 2)
where AO is the spin orbit coupling constant for the free metal ion.
The density of unpaired electrons at the central metal atom was computed
using the equation
45
ri eu = (AIIIP) + ( gll-2 ) + 3/7(gJ.-2) + 0.04
where, a
2
gives information about the nature of the bonding. Complete ionic
bonding is indicated bya
2
value of 1, and a? value of 0.5 suggests complete
covalent bonding. Covalency associated with the bonding of metal ion to the ligand
is measured as 1_a
2
.
2.7.6. Magnetic measurement studies
Magnetic susceptibility studies were carried out at room temperature on a
simple Gouy balance. The studies were made using Hg[Co(NCS)4] as standard, as
suggested by Figgis and Nyholm [IS, 19], and diamagnetic corrections were applied
[20]. The effective magnetic moment was calculated using the equation
/leff = 2.S4 (X'm T)112 BM where
T = absolute temperature
X'm = molar susceptibility corrected for diamagnetism of all atoms
present in the complex using Pascal's constant and that of
the zeolite frame work per unit metal.
2.7.7. Scanning electron microscopy
The morphology of the sample is examined using scanning electron
microscopy. Electron microscopy involves the kinetic energy analysis of electrons
ejected out of a specimen when it is subjected to a beam of electrons. The analysis
is carried out by passing a narrow beam of electrons from a tungsten filament over
the surface of the sample and determining the yield of secondary or back-scattered
radiation as a function of the primary beam. Scanning electron microscopy analysis
of a representative zeolite complex before and after soxhlet extraction was
performed on a Leica Steroscan-440 microscope.
46
2.7.8. Thermo gravimetric analysis
Thermal analysis incorporates those techniques in which some physical
parameters of a given system are determined as a function of temperature. Thermo
gravimetric analysis is carried out to study the thermal stabilities of neat and zeolite
encapsulated complexes. Thermo gravimetry is a technique in which a substance in
a desired environment is heated or cooled at a controlled rate and the weight or
mass of the substance is recorded as a function of temperature. The analysis
provides a quantitative measurement of weight change associated with the thermal
decomposition reaction. Using TG it is possible to explain the thermodynamics and
kinetics of various reactions. Thermo gravimetric analysis was performed on a
Shimadzu TGA-SO / TGAVS.IA du Pont / NETZSCH Simultaneous Thermal
Analysis System at a heating rate of 10C/minute in air/an inert atmosphere in the
temperature range 30-800 0c. About 10mg sample was taken for each analysis in a
platinum crucible hung from one arm of the balance in the instrument. The TG data
were computer processed and plotted, with the % weight against temperature.
2.7.9. Conductance measurement
The molar conductance of the simple complex was determined at room
temperature in methanol (10'3 M) using an 'Elico PR 9S00' conductivity bridge
with a dip type cell and a platinised platinum electrode.
2.8. Catalytic studies
2.8.1. Gas chromatography
Gas chromatographs are essential instruments in catalysis research as they
are used for the separation of molecules. Gas Chromatograph Chernito 8S10 was
used for analysing the reactants and products of the catalytic reactions over zeolite
47
complexes. An SE-30 column was used for separating various components in the
reaction mixture.
The sample is injected into a heating block where it is immediately
vaporised and swept by the carrier gas stream into the column. The solutes are
adsorbed at the head of the column by the stationary phase and then desorbed by
the carrier gas. The adsorption- desorption process occurs repeatedly as the sample
moves towards the column outlet. Each solute travels at its characteristic rate
through the column. The bands separate to a degree that is determined by the
individual partition rates. The solute then enters the detector attached to the column
exit. The peaks that appear on the recorder are characteristic of the different
components. The peak area is proportional to the concentration of the component
in the mixture.
2.8.2. Thin layer chromatography
Glass plates coated with dried and activated silica gel (adsorbent) were used
for thin layer chromatography. The substance to be separated was spotted near one
end of the adsorbed layer and was placed vertically in a jar containing a suitable
solvent or a mixture of solvents. The solvent passes through the adsorbed layer as
per the principle of capillary action. The plate was taken out from the jar and kept
in air. Visualisation was achieved by exposure to iodine vapours.
48
References
1. P.K Dubey and Ramesh Kumar, Ind. 1. Chem. 38 B (1999) 121l.
2 AI. VogeL A Text Book of Practical Organic Chemistry, Longman, London,
1972.
3 KKM. Yusuff, Ph.D. Thesis, University of Kerala, Trivandrum, August 1976.
4 M. Kumar and S. Arabinda Kumar, Asian 1. Chem. 6, 4 (1994) 782.
5 Jayendra Patole, Sabari Dutta, Subhash Padhye, and Ekkehard Sinn, Inorganica
Chimica Acta 318 (2001) 207.
6 Edward H. Yonemoto, Yeong 11 Kim, Russel H. Schmehl, Jim O. V/allin, Ben
A Shoulders, Benny R Richardson, James F. Haw, and Thomas E. Mallouk,
1. Am. Chem. Soc. 116 (1994) 10537.
7 P.G. Menon, Lectures on Catalysis, 4Ft Ann. Meeting, Ind. Acad. Sci., S.
Ramasheshan (Ed.) 1975.
8 Robert 1. Taylor, Russel S. Drago, and James E. George, 1. Am. Chem. Soc.
111 (1989) 6610.
9 N. Herron,Inorg. Chem. 25 (1986) 4714.
10 N. Ulagappan and V. Krishnasamy,Ind. 1. Chem. 35A (1996) 787.
11 A.I. VogeL A Text Book of Quantitative Inorganic Analysis, (4th edn.),
Longman-Green, London, 1978. Rpt. 1986.
12 S. Brunauer, P.H. Emmett, and E.J. Teller, Am. Chem. Soc. 60 (1938) 309.
13 H.G. Hecht, Modern Aspects of Reflectance Spectroscopy, W.W.
Wendlandt, (Ed.) Plenum Press, New York, 1968.
14 S.K Tiwary and S. Vasudevan,Inorg. Chem. 37 (1998) 5239.
15 R W. Frei and J.D. Mc NeiL Diffuse Reflectance Spectroscopy in
Environmental Problem Solving, CRC Press, Cleveland (1973).
16 B.v. Agarwala, Inorg. Chim. Acta. 36 (1979) 209.
17 D. Kivelson and R Neiman, 1. Phys. Chem. 35 (1961) 149.
18 B.N. Figgis and RS. Nyholm, 1. Chem. Soc. (1958) 4190.
19 B.N. Figgis and 1. Lewis in Modern Coordination ChemistlY, J. Lewis and
RC. Wilkins (Eds.) Interscience, New York 1960.
20 R.L. Dutta and A Syamal, Elements oJMagnetochemistry, S. Chand and Co., New
Delhi, 1982.
49
Chapter 3
Zeolite-Y encapsulated complexes of the Schiff bases
)-1 ,2-phenylenediamine,
'-bis(3-pyridylidene )-1 ,4-phenylenediamine,
and salicylaldehyde semicarbazone
Abstract
This chapter, dealing with the synthesis and characterization of zeolite encapsulated Schiff base
complexes, is divided into two sections. The ftrst section gives a comparative study of the
complexes of Fe(III), Co(lI), Ni(II) and Cu(U) with -pyridylidene)-1 ,2-phenylenediamine
and NN '-bis(3 -pyridylidene )-1 A-phenylenediamine. The second section deals with study of Co(II),
Ni(II), and Cu(II) with salicylaldehyde sernicarbazone. Elemental analysis and CHN analysis were
done to understand the stoichiometry of the complex. Si/AI ratio and XRD patterns suggest that
there is no change in the crystalline nature of the zeolite even after encapsulation. Magnetic
susceptibility measurements diffuse reflectance and EPR spectra gave information regarding the
geometry of the complex. The formation of complex inside the zeolite pore was confmned by IR
spectra. The TGIDTG patterns gave qualitative idea regarding the thermal stability of the complex
and the probable composition of the expelled group.
3. Introduction
Schiff bases contain azomethine groups, and the stability of the Schiff base
depends on the strength of the C=N bond, basicity of the imino group and steric
factors. Recent advancements in the coordination chemistry of Schiff base are
essentially concerned with the properties and structures of metal chelates. The
presence of electron releasing or electron withdrawing groups or an aromatic ring
can tune the electronic and magnetic properties of metal complexes of Schiff
bases. The presence of a second functional group with a replaceable hydrogen
atom, preferably a functional group close to the imino group, allows the ligand to
form a fairly stable four, five, six, membered ring on chelation to the metal atom.
The relative stability of such ligands allows distortion in the structure of the
complexes. The extent of distortion depends on the nature of the metal and the
apical and axialligands.
The complexes of Schiff bases are found to have antibacterial, antifungal,
anti-inflammatory, antileukaemic and such other properties [1-3]. These
complexes have wide applications, as homogeneous catalysts for oxidation,
carboxylation and decarboxylation [4,5]. Schiff base complexes also find
applications as heterogeneous catalysts for these reactions. One of the methods for
modifying the homogeneous catalysts into heterogeneous catalysts is the
encapsulation of the complexes inside zeolite cages. It was considered worthwhile
to synthesize the complexes inside zeolite cages and understand the nature of the
complexes inside the cages.
In this chapter, our studies on the zeolite encapsulated Fe(III), Co(II), Ni(II)
and CU(II) with N,N '-bis(3-pyridylidene )-1 ,2-phenylenediamine, N,N '-bis(3-
pyridylidene)-1,4-phenylenediamine and Co(Il), Ni(II), and CU(II) with
salicylaldehyde semicarbazone are presented. For convenience the chapter is
divided into two sections, Section A and Section B. Section A deals with the
51
studies on the Fe(III), Co(II), Ni(II) and Cu(U) complexes of with N,N'-bis(3-
pyridyIidene )-1 ,2-phenylenediamine, and N,N '-bis(3-pyridylidene )-1,4-
phenylenediamine and Section B, deals with the studies on Co(II), Ni(II), and
CU(II) complexes with salicylaldehyde semicarbazone.
Section A
Zeolite encapsulated Fe(ill), Co(II), Ni(II) and CU(II)
N,N' -bis(3-pyridylidene )-1,2-phenylenediarnine and
N,N' -bis(3-pyridylidene )-1,4-phenylenediamine corn plexes
3A.1. Introduction
Schiff bases may form monomeric or dimeric neat complexes with metal
ions. N,N'-Bis(3-pyridylidene)-1,2-phenylenediamine forms both monomeric and
dimeric complexes with Fe, Co, Ni and Cu [6,7]. Complexes of Fe, Co, Ni and
Cu with N,N '-bis(3-pyridylidene )-1 ,2-phenylenediamine (SPO) were encapsulated
in zeolite and it was found that the encapsulated complexes formed only
monomeric species. The experiment was repeated with NN'-bis(3-pyridylidene)-
1,4-phenylenediamine (SPP) encapsulated in zeolite. These encapsulated
complexes were also found to form only monomeric complexes.
H ~ H
~
l __ J N N-
N Ij
Figure 3A.1 Structure of N,N '-bis(3-pyridylidene )-1 ,2-phenylenediamine (SPO)
52
H
Figure 3A.2 Structure of N,N '-bis(3-pyridylidene )-1 ,4-phenylenediamine (SPP)
3A.2. Experimental
3A.2.t. Synthesis of ion exchanged zeolites
Details regarding the synthesis of ion exchanged zeolites are gIven ID
Chapter 2. The metal ions have preference for certain sites in the zeolites. After
metal exchange the metal ions in the zeolite may exist in the form of hexa aqua
complexes, which are localized in the supercage. On dehydration the coordinated
water molecules are removed and the vacant sites are occupied by lattice oxygen.
The metal can migrate to other sites also. Rehydration shifts the cation back to the
supercage. Thus the metal ion can be considered to be in dynamic state, where the
metal migrates to the most favourable position to get the best coordination.
3A.2.2. Synthesis of zeolite encapsulated Fe(ill), Co(ll), Ni(II) and CU(II)
complexes of N;N'-his(3-pyridylidene)-1,2-phenylenediamine, and
N;N' -his(3-pyridylid ene )-1 ,4-phenylened iamine
Metal exchanged zeolite (3.0g) was added to a solution of the ligand in
ethanol, keeping the ligand to metal mole ratio - 2 (slightly higher than 2). The
mixture was refluxed for six hours to ensure complexation. The ligand penetrates
53
through the pores of the zeolite and complexes with the metal ions already present
within the zeolite. The complexed product becomes too large to pass through the
aperture of the supercage and will be retained inside the cages. The resultant mass
was soxhlet extracted with methanol until the extracting solvent became
colourless, ensuring the complete removal of surface species (ligand and
complexes adhering to the surface). The uncomplexed metal ions in the zeolite
and ionisable protons of the ligand were removed by ion exchange with NaCI
solution (0.1 M, 250ml) for 24 hours. It was filtered, washed free of chloride ions
and finally dried at 100C for 2 hours and stored in vacuum over anhydrous
calcium chloride.
3A.3. Characterization techniques
Details regarding the characterization techniques are given in Chapter 2.
3A.4. Results and Discussion
3A.4.1. Ion exchanged zeolite
3A.4.1.1. Chemical analysis
The analytical data of NaY and metal exchanged zeolites are presented in
Table 3A.1. The Si/AI ratio for NaY is found to be 2.4 which corresponds to a unit
cell formula of Na56(AI02)56(Si02)136 [8]. The unit cell formula represents the
composition of a unit cell in the metal exchanged zeolite. The Si/AI ratio remains
the same in all the metal exchanged zeolites, indicating that no destruction of
zeolite framework has occurred during ion exchange by dealumination.
Preservation of the zeolite framework on exchanging with dilute metal salt
solutions has been noted by earlier workers also [9]. The process of ion exchange
can be represented as
Na56
Y
+ XM2+
M"Na56-2" Y + 2x Na ...
54
Table 3A.1 Analytical data of metal exchanged zeolites
Sample % Si %Al %Na %M
....................................................................................................................................................................
NaY 21.76 8.60 7.5
FeY 21.62 8.56 6.2 0.53
CoY 21.71 8.50 5.9 0.25
NiY 21.59 8.55 6.1 0.50
CuY 21.48 8.48 6.6 1.12
PdY 21.50 8.51 6.9 0.164
Table 3A.2 Composition of metal exchanged zeolites
Ion
exchanged
zeolite
NaY
FeY
CoY
NiY
CuY
PdY
Degree of
ion exchange
(%)
8.97
2.67
5.30
10.90
0.97
Unit cell formula
Na56(AlO2)56(Si02)136nH20
Na50.97FeI.676(Al02)56(Si02)136 nH20
Na54.5CoO.749(Al02)S6(Si02)136 nH20
Na53.0 Ni1.5oo(Al0
2
)56(Si0
2
)136 nH20
Na49.88CU3.060(Al02)56(Si02)136 nH20
Na55.45Pdo.270(AI02)56(Si02)136 nH20
55
where x represents the atom fraction ofM
2
+ ions introduced into the zeolite [10].
Strong interaction between the support and the metal leads to high loadings of
metal ions inside the zeolite. The loadings in the metal exchanged zeolites were in
the range 0.5% to 1.12%. The degree of ion exchange was found to increase in the
orderPd < Co < Ni < Fe < eu.
The degree of ion exchange is represented as the percentage of Na + ions
replaced by metal ions from the total amount of sodium which is equivalent to the
aluminium content of the zeolite. The value of degree of ion exchange in various
metal exchanged zeolites used in the present study is comparable to that reported
in the literature [11].
3A.4.1.2. X - ray diffraction patterns
The X-ray diffraction patterns of the zeolite samples NaY, and CoY are
given in Figure 3A.3. The XRD patterns of metal exchanged zeolites are very
similar to those of the parent HY zeolite as well as to those reported in literature
[12]. This indicates that the crystalline structure of the zeolites is preserved even
after the ion exchange. The crystalline structure of zeolite is known to be affected
by dealumination on metal exchange using metal salt solution of concentration >
0.02 M and pH < 4.0 [9]. Therefore in the present study very low concentration of
the metal ion is used. Further, crystalline phases of metal ions were not detected in
any of the patterns. This implies that the metal ions are finely dispersed at the
cation phases of the zeolite rendering them undetectable by XRD.
56
(ii)
Figure 3A.3 XR.D patterns of (1) NaY, (2) CuY
57
3A.4.1.3. Surface area and pore volume
The surface area and pore volume ofHY and metal exchanged zeolites are
detennined by nitrogen adsorption at low temperature and at relative pressures
PfP
o
in the range 0.1 to 0.9. The surface area and pore volume data are given in
Table 3A.3. The values indicate that the introduction of metal ions into the zeolite
lattice by ion exchange causes only marginal reduction in the surface area. This
helps to rule out the possibility of the destruction of the zeolite matrix on ion
exchange. The surface area obtained for the metal exchanged zeolites agrees with
the values reported earlier [13].
Table 3AJ Surface area and pore volume data ofHY and ion exchanged zeolite
, -------------,,-
Sample Surface area Pore volume
m
2
/g ml/g
~ ~ - ~ ~ ~ ~ ~ ~ ~ ~ - - ' ' ' ' ' ' ' ' ' ' - ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' - ' ' ' ' ' ' ' - ' ' ' ' .... __ ... _ .... _ .. _ .... _. __ .......................... ..
HY 546 0.3045
NaY
FeY
CoY
NiY
CuY
PdY
3A.4.1.4. FTIR spectrum
540
529
532
528
530
500
0.3021
0.2982
0.2986
0.2974
0.2983
0.2870
The FTIR. spectrum of NaY is shown in Figure 3A.4. The IR. absorptions in
both the hydroxyl region (wave number beyond 3000 cm-I) and the framework
region (wave number below 1500 cm-I) are listed in Table 3A.4. Framework
vibrations are observed in the spectral range between 1250 cm-
I
and 400cm-
l
.
Symmetric stretching modes, antisymmetric stretching modes, bending modes of
58
T0
4
[(Si/AI)04] tetrahedra, and structure sensitive vibration band (called double
ring vibration band) [14,15] can be seen in Figure 3A.4. The broad absorption at
- 1 DDOcm"1 is attributed to external asymmetric stretching vibration of T0
4
. The
symmetric stretching T-O-T vibration gives rise to bands seen at 630-870cm"l. The
stretching and bending vibrations of water molecules present in the zeolite lattice
can be seen at - 3500 cm"1 and - 1640 cm"1 respectively [16].
Table 3A.4 IR. spectral data of metal exchanged zeolites
Range of wave numbers
"I

cm
1250-980
870-630
550-620
ca. 450
Assignment
asymmetric T -0-T stretching
symmetric T -0-T stretching
structure sensitive double five ring vibration
T -0-T bending mode
- - r--.. -T--.. -T--------r----------r--------r-.. --- -T-- -----T-------,---i
1000 375lJ J500 :i2511 .UlO 2?5IJ :I5W 2IJlC 175(1 !5a1 !25tl lOCC 500
Wil'ir.a.aeor c.-:
Figure 3A.4. FTIR. spectrum of NaY
59
The IR. bands of metal exchanged zeolite appear almost at the same position as
those of the parent zeolite Y. Further, these values are almost the same as the
values reported in earlier studies [14,15,17]. Therefore the zeolite framework
remains unaffected on ion exchange usmg the metal ion solution of low
concentration used in the present study.
3A.4.l.S. Diffuse reflectance spectra
The electronic spectra of metal exchanged zeolites are given in Figure
3A.5. The transition metal ions get coordinated to the oxide ions in the lattice and
to the water molecules present in zeolite. These being weak ligands, the complexes
are expected to be in high spin state. The data indicate octahedral structure for the
species inside the zeolite cages [18].
3A.4.1.6. EPR spectrum
The EPR parameters of copper exchanged zeolites are given in Table 3A.5.
The gll value is higher than the usually reported value, but such high value has
been observed by Sakaguchi and Addison [19] for copper exchanged zeolites. The
high value can be attributed to the tetrahedral distortion of a square planar
chromophore, decreased covalency of metal ligand bond, or due to the increasing
positive charge on the donor atom.
Table 3A.S EPR parameters of copper exchanged zeolites
Ion exchanged zeolite
CuY 2.37 2.08 136.66 66.66
60
(1 )

Wavelength (nm)
...

.DD
I>
"
c:
..
.<l
0
.,
.<l
cC
(2)
8
c:
..
-e
o
1l
cC
0.8
0.6
0.'
0.3
0.2
400 800
Wavelength(nm)
800
D.D
400 60D
'DD
Wavelength (nm)
(4)
(1) FeY
(2) CoY
(3) NiY
(4) CuY
(5) PdY
Figure 3A.S Electronic spectra of ion exchanged zeolites
61
3A.4.2. Zeolite encapsulated complexes
3A.4.2.1. Chemical analysis
The process of complex fonnation inside the cage for divalent metal ions
can be represented as:
+ L - [(MLn)]xNa56-2xY
where L represents the ligand coordinated to the metal center [10].
The analytical data of the zeolite encapsulated complexes are given in Table 3A.6.
Analytical data show that the Si/AI ratio of the zeolite complex is 2.4 as in the case
of metal exchanged zeolites. This again is an indication of the zeolite framework
remaining unaffected on encapsulation. Similar observation with regard to the
Si/AI ratio has been made by earlier workers [20,21]. The ligands, SPO and SPP,
form complexes in the super cages probably by breaking the metal-lattice oxygen
bonds [22].
Table 3A.6 Analytical data
Sample % Si %AI %Na %C %M
FeYSPO 18.64 7.60 5.40 3.60 0.50
CoYSPO 18.46 7.48 5.36 l.35 0.23
NiYSPO 18.54 7.63 5.38 3.00 0.47
CuYSPO 18.71 7.67 5.41 5.60 0.93
FeYSPP 18.79 7.83 5.60 3.53 0.51
CoYSPP 18.79 7.74 5.90 l.22 0.22
NiYSPP 18.90 7.91 6.00 l.81 0.30
CuYSPP 18.93 7.82 5.85 5.45 0.89
_ ............................_--_ ....
------_ ..........._--.........._-
62
The positive charge of the encapsulated complex will be compensated through the
interaction of negatively charged oxide ions of zeolite matrix. The loadings of the
metal in the ion exchanged zeolite and the zeolite encapsulated complexes are
quite close in value to each other which suggests that almost all the metal ions
present in the zeolite lattice get complexed and the uncomplexed metal ions are
removed during back exchange. The ligand to metal mole ratio is 2 in neat
monomeric complexes of Co(II), Ni(II) and CU(II) with SPO ligands [6,7]. Though
the same value is expected in zeolite encapsulated complexes, the actual value is
observed to be slightly lower than 2. This could be because some metal ions get
trapped in the cavities of the zeolite and the metal complexes that surround them
prevent ion exchange with sodium ions.
3A.4.2.2. Scanning electron micrographs
Scanning electron micrographs of zeolite encapsulated species before and
after soxhlet extraction are shown in Figure 3A.6. In the SEM taken before soxhlet
extraction, the surface is crowded with particles. In the SEM taken after soxhlet
extraction, the surface is clean, indicating the removal of the surface species
during soxhlet extraction. The complexes formed inside the zeolite cages remain
trapped and are not able to come out. Similar reports of clear zeolite surface in the
case of phthalocyanine and salen are seen in the literature [23,24].
3A.4.2.3. X-ray diffraction patterns
X-ray diffraction patterns of both the SPO and SPP complexes are given in
Figure 3A.7 and Figure3A.8. The XRD patterns of the complexes were compared
with those of NaY and ion exchanged zeolite. The XRD patterns are found to be
similar to those of the corresponding metal exchanged zeolite and the parent
zeolite.
63
(i)
(ii)
Figure 3A.6 SEM
i) Before soxhlet extraction
ii) After soxhlet extraction
64
5 10 15 20 25 30 35 40
Figure 3A. 7 XRD patterns of SPO complexes
i) FeYSPO, ii) CoYSPO, iii) NiYSPO, iv) CuYSPO
65
5 10 15 20 25 30
Figure 3A. 8 XRD patterns of SPP complexes
i) FeYSPP, ii) CoYSPP, iii) NiYSPP, iv) CuYSPP
66
This indicates that the zeolite structure has not been damaged even after the
synthesis of metal complexes in the cavities. It has been reported that the presence
of the large encapsulated phthalocyanine complex does not make the zeolite
network more fragile [25].
3A.4.2.4. Surface area and pore volume
The surface area and pore volume of encapsulated SPO and SPP complexes
of Fe, Co, Ni and eu were measured by nitrogen adsorption. The surface area of
zeolite is mainly internal surface area, and so it should decrease when there is
blocking of the pores. The data (Table 3A.7 and Table 3A.8) reveal that the
surface area values are lower in the case of zeolite encapsulated complexes than
those of the corresponding metal exchanged zeolites. The lowering of surface area
and pore volume can be attributed to the filling of the zeolite pores with metal
complexes. Such lowering of surface area and pore volume on encapsulation has
been reported by earlier workers [11,26,27]. The loss in surface area varies
between 68.3% and 73.3% for the encapsulated SPO and SPP complexes and the
loss in pore volume varies between 74.6% and 76.6% for the encapsulated SPO
and SPP complexes. This provides strong evidence of the encapsulation of the
complex within the zeolite pore.
Table 3A.7 Surface area and pore volume data of complexes of SPO Ligand
- - ' - ' . ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' - ' ' ' - ' ' ' ' ' ' ' ' ' ' ' ' ' - ' ' ' . ' ' ' ' ' ' ' ' ' ~ ' ' ' ' ' ' ' ' ' ' ' ' ' , , ' ' ' - , ' ' , , ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' - ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' ' '
Surface area (m /g) Pore volume (ml/g)
.................................. nu ..............................................................................................................................
Sample MY MYSPP % loss MY MYSPP % loss
FeY 529 170 68.3 0.2982 0.0758 74.6
CoY 532 155 70.9 0.2966 0.0706 76.2
NiY 528 152 71.2 0.2974 0.0701 76.2
CuY 530 147 71.7 0.2983 0.0691 76.4
67
Table 3A.8 Surface area and pore volume data of complexes of SPP ligand
~ ' ' ' ' ' ' ~ - ' ~ ' ' ~ ' - - ~ - - - ' ' - ~ - - - - ~ - ' ~ ' - - - ' ' ' ' ' ~ ' - - ' ' ' - ' ' ' ' - - ' - ' ' ' ' ' ' ' - - - - - - ' ' ' ' ' ' '
Surface area (m /g) Pore volume (ml/g)
Sample MY MYSPP % loss MY MYSPP % loss
FeY 529 162 69.8 0.2982 0.0710 76.2
CoY 532 150 7l.8 0.2966 0.0700 76.4
NiY 528 145 72.5 0.2944 0.0690 76.6
CuY 530 139 73.3 0.2930 0.0685 76.6-
' - ~ ' - - - - - ...... - ...--------,----..............--...-........-,,--...-----......-, ... -..........
lA.4.2.S. Magnetic moment
The magnetic moments were measmed at room tempeIature usmg Gouy
method. This method gives only qualitative idea regarding ma.gnetic Illll[)ment due
to high diamagnetic contribution from zeolite framework, paIramagnetic
contribution from the trace iron present in the zeolite lattice as impurity, and also
because of the very low concentration of the complex within the zeolite.
Table 3A.9 Magnetic moments of SPO and SPP complexes
- - , - - - - - - - - - - - - ~ .. _--------,-
SampJe Magnetic moment (BM)
Gouy method EPR.
FeYSPO 5.94
CoYSPO 5.10
NiYSPO 3.00
CuYSPO l.84 l.87
FeYSPP 5.90
CoYSPP 5.15
NiYSPP 3.l0
CuYSPP l.86 l.87
68
However the susceptibility of NaY was measured and its contribution to the
magnetic moment of the encapsulated complex per mole of metal ion was
calculated. The diamagnetic contribution of the ligand molecule was also taken
mto account.
The room temperature magnetic moments of the SPO and SPP complexes
are given m Table 3A.9. The values indicate that all the complexes are
paramagnetic. The magnetic moment of FeYSPO and FeYSPP are 5.94 and 5.90
BM. These values are closer to those of high spin Fe(III) complexes [28],
suggesting a high spin octahedral geometry for the Fe(III) encapsulated complex.
The magnetic moment is not of much use in identifying the geometry in the case of
neat Fe(III) complexes, but is useful in assigning geometry to neat Co(lI)
complexes [29]. Octahedral CO(Il) complexes normally have magnetic moments
m the range 4.80 to 5.20 BM which is higher than spin only values of 3.89 BM.
This difference from the spin only value is due to the considerable orbital
contribution [29,30]. The magnetic moments of 5.10 and 5.15 BM for CoYSPO
and Co YSPP indicate an octahedral structure. The simple CO(lI) SPO monomeric
complex is reported to have square planar geometry [6]. In the case of zeolite
encapsulated complexes, the oxide ions of the zeolite lattice or water molecules
can also participate in the coordination to the metal ion resulting in an octahedral
geometry.
The magnetic moment of NiYSPO is 3.00 BM and that ofNiYSPP is 3.10
BM. The value suggests an octahedral geometry for the complex as Ni(II)
octahedral complexes usually have magnetic moments in the range 2.90-3.30 BM.
The neat complex of SPO with Ni(lI) is reported to have square planar geometry
[6]. The change in geometry can be attributed, as in the case of zeolite
encapsulated cobalt(II) complex to the interaction of the zeolite framework on the
Ni(Il) complex. The magnetic moment values are higher than the spin only values
of 2.83 BM, suggesting some orbital contribution to magnetic moments which
69
may arise due to mixing of the ground state with the first excited triplet state 3T 2g
having electronic configuration h
g
S
eg
3
.
The magnetic moments of CU(II) complexes are generally in the range of
1.7-2.2 BM [30]. CuYSPO and CuYSPP have magnetic moments l.84 and 1.86
BM respectively. Cu(II) square planar complexes show magnetic moments near the
lower limits. The low value in the present case suggests square planar structure for
the two CU(II) complexes.
3A.4.2.6. Diffuse reflectance spectra
Electronic spectra in the diffuse reflectance mode are preferred over the
absorbance mode for zeolite compounds since the radiations scattered by zeolites
interfere with absorption due to electronic transition. The electronic spectral data
of encapsulated complexes and their tentative assignments are given in Table
3A.10 and Table 3AII, and the spectra are given in Figure3A 9 and Figure3AI O.
The electronic spectra of Fe(III) complexes of SPO and SPP show a broad band
with a maximum around 22000cm"l. This corresponds to a combination of 4T Ig -
6
A1g
, 4
T2g
_ 6 AIg. 4E g _ 6 A
1g
spin forbidden transitions in octahedral symmetry
[31]. The band at 18180cm"1 seen in the case of Fe YSPP may be due to spin
forbidden d-d transition in high spin Fe(III). The bands at 29400cm"1 and 33300
cm
1
in the spectra of FeYSPO and FeYSPP (respectively) can be due to charge
transfer transition [32].
Co(II) (d
7
system) complexes are expected to show three spm allowed
transitions in octahedral geometry corresponding to transition ~ 2g (P) - 4T Ig(F),
4
A2g
(F) _ ~ 1 g { F , and 4
TIg
(P) - 4
TIg
(F). The band V2 is a two electron transfer
from t2/ to hg
3
eg
4
and therefore would be of very low intensity [30]. In the case of
CoYSPO and CoYSPP the bands at 16390cm"land 15150cm"1 correspond to
4A2g(F) - 4T Ig(F) transition and the bands at I 9600cm-
1
and 21600cm"1 can be
assigned to 4
T1
g{P) - 4T
Ig
(F) transition in octahedral geometry [6].
70
Ni(II) complexes are reported to exhibit three spin allowed transitions in
the regions 8000cm-
I
-29000cm-
1
in cubic field. In the case ofNiYSPO all the three
expected transitions are seen. One is observed at 12500cm -I corresponding to
lT
2
g(F) _ 3 A
2g
(F), another at 18000cm-
1
due to 3T Ig(F) - 3 A
2g
(F), and the third
one is observed at 28000cm-
l
, due to 3T Ig{P) - 3 A
2g
(F) transitions. In the case of
NiYSPP, of the three expected transitions, only two are seen, one at 14000cm-
1
corresponding to 3T Ig(F) - 3 A2g(F) transition and the other at 25000cm -I due to
)T1g(P) - 3 A2g(F) transitions [30 ].
Copper(II) complexes are susceptible to Jahn-Teller distortion and hence
they do not form regular octahedral complexes. Tetragonally distorted octahedral
complexes, which finally end up with square planar structures are usually formed.
Table 3A.I0 Electronic spectral data of SPO complexes
Sample Abs. max. Assignment
.-... --......... --.............. -... -............ ~ ~ . ~ ~ ........ -................. -.-..... -......... -.. -... _ .... _._ ...... _ ..... _ .... -....... -
FeYSPO
CoYSPO
NiYSPO
CuYSPO
22000
29400
16390
19600
29940
12500
18000
28000
30150
13160
29000
4 A2g(F) _ 4T 1 g(F)
4
T1
g(P) _ 4 Tlg(F)
Charge transfer transition
3
T
2g(F) _ 3 A2g(F)
3
T1
g{F) _ 3 A2g{F)
3
TI
g{P) _ 3 A2g(F)
Charge transfer transition
d-d transition
Charge transfer transition
----, ----
71
025
.,
'"""'-
0
c
.e 020
0
..
IJ
c(
015
010
0.05
! ~ .
0.24 ..
022
8
c
.e 0.20
o
..
IJ
c(
0.18
018
'\
400
400
I FeYSPO I
600
Wavelength (nm)
,
\
I NiYSPO I
\,,"
600
Wavelength (nm)
BOO
800
B
c:
ca

o
~
I CoYSPO I
\/-..,
" ... '"
~ .
400 600 BOO
Wavelength (nm)
1.5 . - - ~ - , - - ~ - , - - ~ - , - - ~ - - , - - . - - - - ,
I CuYSPO I
0.5 - - ~ - - - - - - ~ - - - - - - ~ - - - r - -
400 600
Wavelength(nm)
soo
Figure 3A.9 Electronic spectra of SPO complexes
72
Table 3A.ll Electronic spectral data of SPP complexes
Sample Abs. max. Assignment
_. ________________ _____________________________ _
FeYSPP
CoYSPP
NiYSPP
CuYSpp
18180
22000
33300
15150
21600
29500
14000
25000
29850
12500
29325
d-d transition
4 6
T
1g
- A
1g
6
A1g
4 6
Eg- A1g
Charge transfer transition
4 A2g(F) _ 4T 1 g(F)
4
T1g
(P) _ 4
T1g
(F)
Charge transfer transition
3
T
1 g(F) _
3
A2g(F)
3T
1g
(P) _ 3 A2g(F)
Charge transfer transition
d-d transition
Charge transfer transition
Many CU(II) complexes give a broad band in the region 11000-16000cm-
1
resulting in the blue or green colour. The bands at 13160cm-
1
and 12500cm-
1
in
the spectra of the complexes Cu YSPO and Cu YSPP correspond to d-d
transitions. Magnetic susceptibility data suggest square planar geometry for the
encapsulated complex. The structures of neat Cu(n) complexes of SPO and SPP
also exhibit bands in the region 13000-14000cm-
1
. The bands at 29000cm-
1
and
29325cm-
t
in the electronic spectra of the encapsulated complexes are due to
charge transfer transitions.
73
0.8
0.6
Cl
0
C
III

0.4 0
en

c(
0.2
081
060
8 0.511
C
III
,
0.58
0.57
0.56
200
IFeYSppl
_.,,--..
400 600 800
Wavelength (nm)
I NiYSppl
\/\"".''' ,
! '.
. .'
,- '.i
0.55 __
400 600 800
Wavelength (nm)
1.2
1.0
0.8
G>
<)
c:
III
0.6
-e
0
..
.Q

0.4
0.2
0.0
0.48
0.46
0.44
ID
<)
c:
III
0.42
-e
0
..
.Q

0.40
0.38
0.36
ii,
1\

,.,
I'
i
\
i
i
i
I
I
1
\
\
'----- --
-----
-----
200 400 600 800
Wavelength (nm)
(\i'\
\
ICuYSppl
!"
\
i
!
\, ....
...
"
,
"-
"-
j ....
.,
\
'-.
-....,.
-.
400 600 800
Wavelength (nm)
Figure 3A.I0 Electronic spectra of SPP complexes
74
3A.4.2.7. FTm spectra
The IR spectra of ligand SPO and zeolite encapsulated SPO complexes are
given in Figure 3A.11, and those of ligand SPP and zeolite encapsulated SPP
complexes are given in Figure 3A.12. The IR. absorption frequencies of SPO
ligand and encapsulated SPO complexes are presented in Table 3A.12, and those
of SPP and its complexes are presented in Table 3A.13. The strong band at
1581cm-
1
in the spectrum of the ligand SPO is shifted to the lower energy region
1560-l545cm-
1
in the spectrum of the complexes, indicating coordination of azo-
methine nitrogen. Some of the bands of the ligand are masked by strong absorption
of the zeolite support. Some new bands seen in the spectrum of the encapsulated
complex are those of the zeolite support. The OH stretching and bending vibration
of water is seen as a broad band at 3450-3470cm-
l
and 1630-1640cm-
l
in the
spectra of all the encapsulated complexes. Bands around 1030cm-
l
and 700cm-
1
may be attributed to the rocking and wagging modes of water.
The band at 1610cm-
l
corresponds to uC=N stretching vibration of the
ligand SPP. This band is seen between 1540 and 1560cm-
l
in all the complexes
mdicating coordination through nitrogen. As in the case of SPO complexes, the
OH stretching vibration of water is seen as a broad band in the range 3440-
3470cm-
1
and bending vibration between 1635 and 1640cm-
l
in the spectra of all
the encapsulated complexes. The rocking and wagging modes of water are also
seen in the spectra of the complexes.
75
Table 3A.12 Infrared spectral data of SPO complexes
--_ .... '"""-, .........--.... _ ......__ ............................................_-_ .......__ .. __ ............_ ...__ .......................................-....................................................................................................._ ..
SPO FeYSPO CoYSPO NiYSPO CuYSPO
cm
-1
cm
-1
cm
-1
cm
-1
cm
-1
---_ ...... --_ .._--_ .... __ .... _ .. __ .... _ ............
470 469 464
536 567
623 630 640 635 640
703 700 700 720 730
747
817
957
1023 1025 1040 1050
1114
1126
1317
1432
1538
1581 1560 1545 1550 1550
1640 1640 1641 1640
1925
2375 2375 2368 2368
2789
3103
3439 3470 3480 3474 3480
---_ .. __ ..._.
76
Wavenumbcr (errr')
Figure 3A.ll IR spectra of SPO complexes
i) SPO ii) FeYSPO, iii) CoYSPO, iv) NiYSPO, v) CuYSPO
77
Table 3A.13 Infrared spectral data of SPP complexes
,,"'''''------,.._ ....._ .....__ ......_-_ .................._--------_ ......_ .............._--_ ..._---...-------------...-........-......--.....-..., ... ~ - - - - - .....,
Spp
FeYSPP CoYSPP NiYSPP CuYSPP
cm
01
cm
ol
cm
ol
cm
-I
cm-
I
0 _______ 0 ______
442
578
625 615 620 607 610
700 723 730 723 726
752
900
1027 1040 1045 1043 1044
1111 1100 1105
1176
1294
1401
1514
1610 1540 1540 1560 1545
1635 1640 1640 1640
2339 2368 2375 2375
2371
3202
3390
3460 3461 3460 3467
--""' ............................., .....", ..'.,............, ...., .. ", ......, ....." ............, .........._ .... ' .....,.,.. '.,.,.......,.,.,.......' ......., .................................................................................................._ ................._ ...................._ .................................."" ............. " .........................................." .....",............, ......
78
i
1500 toOQ ,.a'" sao
Figure 3A.12 IR. spectra of SPP complexes
i) SPP, ii) FeYSPP, iii) CoYSPP, iii) NiYSPP, iv) CuYSPP
79
3A.4.2.8. EPR spectra
The powder spectra of the zeolite encapsulated CU(IJ) complexes of SPO
and SPP were recorded at room temperature and at liquid nitrogen temperature.
Glass spectra of neat CU(JI) complexes of SPO and SPP were also recorded at
liquid nitrogen temperature. EPR parameters were determined without computer
simulation in the present study. However these parameters provide a qualitative
idea about the nature of coordination in the encapsulated complexes. EPR spectra
of neat and encapsulated SPO and SPP complexes are shown in Figure3A.13 and
Figure3A.14. The EPR parameters, unpaired electron density and magnetic
moment were calculated from EPR spectral data, and the results are given in Table
3A.14 and Table 3A.15. The well defined hyperfine features similar to that in
dilute solution in the case of encapsulated complexes indicate the encapsulation of
monomeric complexes in the zeolite lattice. The gll values of Cu YSPO and
CuYSPP are 2.35 and 2.36 respectively. These are higher than the expected
values. The reasons for such high gll values have been discussed by Sakaguchi and
Addison [19]. The high g value obtained in the present case may be due to the
distortion of structure caused by the zeolite ligand and the zeolite oxygen atom on
coordination. The trend gll > f4 > 2.0023 observed for this complex shows that the
unpaired electron is delocalized in the dx
2
-l orbital [33]. This is typical of
tetragonal distortion of CU(JI) environment which finally leads to a square planar
structure.
The ratio gill All is 133cm and 125cm respectively for the encapsulated SPO
and SPP complexes. These values come within the range 105-13 5cm reported for
square planar complexes and suggest square planar structure for the encapsulated
CuYSPO and CuYSPP complexes. Hyperfine splitting was clearly seen in the
spectrum of neat copper SPO complex while it was not observed in the spectrum
of neat copper SPP complex even at liquid nitrogen temperature. The broad EPR
spectrum in the case of neat copper SPP complex can be due to the nearest
80
neighbour's spin-spin interaction. The spectra of the copper complexes at room
temperature showed no splitting at all.
Table 3A.14 EPR parameters of SPO complexes of copper
EPR parameters CuSPO CuYSPO
._ ..-.. ..__ .. _-_ ..-.... _---_ ...._----
gll gll 2.35
All
225 x 10
4
cm-\ 179 x 10
4
cm-\
g"/AII
107 133cm
~
2.09 2.06
A.l
30.88 x 10
4
cm-\ 59.36 x 10
4
cm-\
a
2
0.9105 0.9132
~ f f
1.9 1.87
P -0.727 -0.6846
--,,-_ ..._ ......._,--,- --,-,_ ..._ ..........-
Table 3A.lS EPR parameters ofSPP complexes of copper
EPR parameters CuSPP CuYSPP
gll
2.4 2.36
All
150x 10
4
cm-\ 188 x 10
4
cm-\
gll IAII 130cm 125cm
g.l
2.04 2.04
sA.l
20 x 10
4
cm-\ 29.1 x 10
4
cm-\
a
2
0.874 0.874
~ f f
1.87 1.87
P -0.5463 -0.5463
81
( iii)
(ii)
(i)
3000
Figure 3A.13 EPR spectra of SPO complexes
i) CuYSPO at RT, ii) CuSPO at LNT, iii) CuYSPO at LNT
82
(ii)
\
/
(i)
Figure 3A.14 EPR spectra of SPP complexes
i) CuYSPP at LNT, ii) CuSPP at LNT
83
The density of unpaired electrons at the metal atom and the magnetic
moments of the neat and encapsulated copper complexes of SPO and SPP were
calculated from EPR parameters. The in plane covalence parameter, a
2
values of
0.9105,0.9132,0.874, and 0.874 for the neat copper SPO, CuYSPO, neat copper
SPP, and CuYSPP respectively indicate an ionic environment for Cu
2
+ ions. The
magnetic moments (l.87 BM for the CuYSPO and CuYSPP complexes) are in
agreement with the values obtained from magnetic susceptibility measurements at
room temperature. All these are suggestive of square planar structure for the
copper encapsulated complexes.
3A.4.2.9. TG analysis
Thermo gravimetric analysis provides a quantitative measurement of any
weight change associated with a transition. The thermal behavior of SPO and SPP
complexes have been studied, and the thermal analytical data for the complexes
are presented in Table 3A.16. The TG curves of SPO complexes are given in
Figure 3A.15 and those of SPP complexes in Figure 3A.16. Generally, well-
defined TG patterns are seen in the case of neat complexes during the removal of
ligand moieties. Such well-defined patterns are not seen in the case of zeolite
encapsulated complexes. Because of the low concentration of metal complexes in
the zeolite cage, a correlation of mass loss with expelled ligand moieties was not
attempted in the present study. However, the TG curves help to get an idea about
the approximate temperature of decomposition. The mass loss around 30-110 QC in
the TG of both the SPO and the SPP series corresponds to the expulsion of
intrazeolite free water molecules. The mass loss in the temperature range 110-500
C corresponds to the removal of coordinated water and the partial decomposition
of the complex and the mass loss corresponding to the temperature range 500-800
C can be due to the decomposition of the complex and partial decomposition of
the zeolite. The IR spectra of the residue after analysis indicate that the
encapsulated complex has completely decomposed. According to reports obtained
84
from literature, the stability of the complexes is believed to be enhanced on
heterogenising them, especially on encapsulating them in zeolite pores [34].
102
lOO
98
98
'I 9-
1:
01
jj 92

90
88
88
I --FeYSPO
o lOO 200 :100 000 500 BOO 700 800 900
. .
TemperalUre ( C)
!--NiYSPO
lOO roll 100 _00 Sltl' Wf) 700 "GO wo
r..,npualUr. (,}C)
100
QS
--CoYSPO
4 100 200 -00 500 11" lOll 800
rempora1Ure (QC)
o 100 21> lOO -00 $(l(- 700 wo
(e}
Figure 3A. 15 TG/ DTG curves of SPO complexes.
85
102,-------------------------------
lOO
98
98
90
88
88
1
,
\
I --F.YSPP
lOO 200 :JOO .00 SOD 800 700 aDo 900
T .m p. 1'31.". (0C)
I --tIIYSPP
I
\
I.

rl

o QI,Q !lOO ?QQ 3QQ
(,;C)
'$.
2!
if

lOO
\
I --CoYSPP
9a
),
96
1\
,.
92
\\.,
\<..-..
80
'\""
'-.
ae
........ --
-
o 100 l!OO :It>O _ SOO 800 700 800 900
T >!m!,erntre (C)
98
88
o .00 800 $\'1
'
VOO
r..,..p .... 1IJr. (OC)
Figure lA. 16 TGIDTG curves of SPP complexes.
86
Table 3A.16 Thermal decomposition data of SPO and SPP complexes
........................................................""-"""'-...... .. --. .... ..................--"" ... .............................................
Sample Temp. Range % Mass loss
TG(C) from TG
............................................................. __ ...... u ........................ __ .......... _ .......... ____ ...... " ......... _ ......................
FeYSPO 30-110 5.75
110-500 8.23
500-800 0.11
CoYSPO 30-110 7.61
110-500 10.5
500-800 1.53
NiYSPO 30-110 4.52
110-500 6.38
500-800 0.59
CuYSPO 30-110 4.87
110-500 5.85
500-800 0.54
FeYSPP 30-110 5.52
110-500 6.83
500-800 0.70
CoYSPP 30-110 4.86
110-500 6.90
500-800 0.48
NiYSPP 30-110 5.37
110-500 6.32
500-800 0.23
CuYSPP 30-110 5.33
110-500 5.79
500-800 0.29
--...... ~ .......---......-.. ............-
87
Section B
Zeolite encapsulated CO(Il), Ni(ll) and eU(Il) complexes of
salicylaldehyde semicarbazone
3B.1. Introduction
Schiff base complexes with oxygen coordination have both structural and
biological significance. These include Schiff bases derived from semicarbazides,
aminophenols, amino acids etc. Semicarbazones act as monodentatel bidentate
ligands depending on the pH of the reacting medium, or as a tridentate ligand
depending on the aldehyde residue attached to semicarbazide moiety. Their
chemistry is interesting as only p-nitrogen coordinates to the metal atom, while the
Cl nitrogen remains uncoordinated. The oxygen atom shows a tendency to fonn
strong covalent bonds with the metal atom.
H
~ / N n N H
Unu 0
OH
Figure 3B.1. Structure of SALSC ligand
Neat complexes with salicylaldehyde semicarbazone have been reported
[35]. Neat copper SALSC complexes have been reported to show superoxide
dismutase (SOD) activity, and it can be increased by the addition of heterocyclic
bases. Studies on the encapsulated Co, Ni, and Cu complexes with salicylaldehyde
semicarbazone were undertaken to know whether any structural and other changes
occur during encapsulation. The results of the studies are presented in this section.
88
3B.2. Experimental
3B.2.1. Synthesis of zeolite encapsulated Co, Ni and Cu complexes of
salicylaldehyde semicarbazone
Metal exchanged zeolite (3.0 g) was mixed with the ligand salicylaldehyde
semicarbazone in such a way that the ligand to metal mole ratio is 1. However a
slight excess of the ligand was used. This was ground well, and heated in a fused
tube (ampule) for 6 hours at 200C and cooled to room temperature. The super
cage entrance is large enough to access the cobalt, nickel and copper sites. The
product was soxhlet extracted first with methanol to remove unreacted ligand, and
then with chloroform to remove any surface species. The uncomplexed metal ions
were removed by sodium exchange, made free of chloride ions, dried at 100C for
2 hours and stored in vacuum desicator over anhydrous calcium chloride.
3B.3. Characterization techniques
Details of the analytical methods of encapsulated SALSC complexes and
other characterization techniques are given in Chapter 2.
3B.4. Results and Discussion
3B.4.1. Chemical analysis
The metal percentage was determined using AAS, and the analytical data
are presented in Table 3B.1. The Si/AI ratio of NaY is retained in the encapsulated
complexes, indicating the absence of dealumination and the preservation of crystal
structure. The metal to ligand ratio is found to be 1: 1 in neat complexes [35]. The
same ratio is expected in the case of encapsulated complexes also. But the
percentage of metal is slightly higher than that required for 1: 1 complexes. This
shows that all the metal ions are not removed by the back exchange ofNa'" ions.
89
The small percentage of metal that remains uncomplexed may be metal
Ions trapped in the cage by the complex formed around it. These traces of
uncomplexed metal ions are unlikely to cause any serious interference in the
behaviour of the encapsulated complexes. Similar observations have been made in
the case of other zeolite encapsulated complexes by earlier workers also [23].
Table 3B.t Analytical data
CoYSALSC 18.75 7.64 5.41 0.25 0.30
NiYSALSC 18.69 7.59 5.85 0.49 0.69
CuYSALSC 18.80 5.79 0.95 l.21
3B.4.2. X-ray diffraction studies
X-ray diffraction patterns of Co, Ni and Cu encapsulated SALSC
complexes and NaY are presented in Figure 3B.2. The diffraction patterns of the
encapsulated complex were more or less similar to those of NaY which suggests
that there is no loss in crystallinity on encapsulation. This again confirms that the
complexes are formed inside the large cavities, and the formation of the complex
does not affect the zeolite lattice.
90
5 10 15 20 25 30 35
Figure 3B.2 XRD patterns of SALSC complexes
i) CoYSALSC, ii) NiYSALSC, iii) CuYSALSC
3B.4.3. Surface area and pore volume
(iii)
(ii)
(i)
40
Surface area and pore volume values of the SALSC complexes are
presented in Table 3B.2. The surface area analyses provide good evidence for
encapsulation, as in the case of SPO and SPP complexes. This decrease in surface
area has been found to be a strong indication of encapsulation [26].
91
Table 3B.2 Surface area and pore volume
- - - - - - - - - ~ - - - - - - - - - - - 1 - - - - - - - - - - - - .-"._._-...
Sample Surface area (m /g) Pore volume (ml/g)
CoY
NiY
CuY
--_._----------------------------------------
MY MYSALSC % loss MY MYSALSC % loss
532
528
530
231
235
240
56.4 0_2986
55_5 0.2974
54_7 0.2983
0.1313
0.1324
0.1331
56.03
55.48
55.38
3B.4.4. Magnetic moment
Magnetic moments of Co, Ni and Cu complexes were calculated, as in the
case of the encapsulated complexes of SPO and SPP. The magnetic moments are
presented in Table 3B.3. The magnetic moments of CoYSALSC and NiYSALSC
complexes nearly agree with the reported values of neat Co(lI) and Ni(II)
complexes ofSALSC (4.93 BM and 3.35 BM respectively), which are suggested
to have an octahedral geometry [36]. It may therefore be inferred that that the
encapsulated complexes also have an octahedral geometry.
Table 3B.3 Magnetic moment
Sample Magnetic moment
(BM)
Gouy method EPR
CoYSALSC 4.85
NiYSALSC 3.16
CuYSALSC 1.78 1.85
92
Neat CU(II) complex is reported to have a value of 1.75 BM [37]. The
encapsulated CU(II) has a value of 1.78 BM by Gouy method and a value of 1.85
BM from EPR measurements. The low value suggests a square planar structure for
the CU(II) complex. Thus the magnetic moments of the neat and the encapsulated
complexes agree with each other.
38.4.5. Diffuse reflectance spectra
The electronio spectra of encapsulated SALSC complexes of CO(ll), Ni(U)
and Cu(II) were recorded in the diffuse reflectance mode . The electronic spectra
of SALSC complexes are presented in Figure 3B.3 and the electronic spectral data
are presented in Table 3B.4. All the three SALSC complexes show strong ligand to
metal charge transfer bands in the region 33000-33560cm-
1
[31].
The electronic spectra of Co YSALSC shows two bands at 15600cm-
1
and
18200cm-
1
attributable to 4A2g(F) - 4TIg(F) (V2), and ~ l g P ) - 4TI g(F) (V3) in
octahedral field [29,31]. All the three transitions corresponding to VI, V2 and V3 in
octahedral geometry for CO(ll) complex have been reported by Maurya et al.
while a tetrahedral structure has been reported by Ray for the same neat complex
[37].
The analysis of the spectrum of NiYSALSC reveals an octahedral
geometry for the metal ion. Of the three spin allowed transitions in cubic field,
only two are observed. The reflectance spectrum exhibits bands at 18000cm-
1
and
25640cm-
1
attributable to 3T Ig{F)- 3 A
2g
(F) and 3T Ig(P) _ 3 A2g(F) transitions
[31]. These bands correspond to the V2 and V3 transitions in octahedral geometry. In
the case of neat complex of SALSC all the three spin allowed transitions are
observed in the regions 8500cm-
l
, 14500cm-
1
and 25000cm-
1
respectively. These
transitions were assigned to 3
T
2g{F) _ 3 A
2g
(F), 3
T
Ig(F) _ 3 A2g(F) and 3
T
Ig(P) _
3 A2g(F) transitions respectively, characteristic of octahedral geometry [36].
93
Table 3B.4 Electronic spectral data of salicylaldehyde semicarbazone complexes
Sample Abs. max. Assignment
(cm-I)
15600
4
A2g
(F) __ 4
TIg
(F)
CoYSALSC 18200 4T
Ig
(P) -- 4TIsCF)
33330 Charge transfer transition
18000
3T
Ig
(F) __ 3 A2g(F)
NiYSALSC 25640 3T Ig(P) -- 3 A
2g
(F)
33560 Charge transfer transition
12000 d-d transition
CuYSALSC 26525 Charge transfer transition
33000 Charge transfer transition
CuYSALSC complex exhibits a broad band at 12000cm-
1
assignable to d-d
transition. The magnetic moment of 1.78 BM supports a square planar structure for
CuYSALSC. The geometry of neat CU(IJ) complex ofSALSC has been reported as
octahedral but, Jayendra Patole's crystal structure studies reveal a distorted square
planar geometry for the copper (n) complex [35].
94
1.2
I CoYSALSC I
1.0
8 O.B
c:
t'II
.0
I-
0
1l 0.6
<{
0.4
0.2
0.0 1....-........ _-'-_ .......... _--'-_ ......... _---1
Q)
o
c:
400 600 BOO
Wavelength (nm)
3.0 "T"""------------.
I CuYSALSC I
2.5
2.0
t'II 1.5
~
o
Cl)
.0
<{
1.0
0.5
0.0 . . L . L . . . r . r . ~
400 600 BOO
Wavelength (nm)
I NiYSALSC I
0.40
0.35
Q)
(J
c:
t'II
.0
....
0
Cl)
.0
<{
0.30
0.25
400 600 BOO
Wavelength (nm)
Figure 38.3 Electronic spectra of
SALSC complexes
95
3B.4.6. FTffi spectra
The ligand SALSC contains three donor atoms, namely, the oxygen of the
phenolic hydroxyl, the nitrogen of the azo-methine group, and the oxygen of the
carbonyl group. The IR spectra of the ligand SALSC and the encapsulated
complexes are given in Figure 3B.4, and Table 3B.5 presents the spectral data.
Table 3B.S IR spectral data of SALSC complexes
SALSC CoYSALSC NiYSALSC CuYSALSC
cm
-1
cm
-1
cm
-1
cm
-1
459
561 570 570 570
611 616 609
679
750 771 774 724
894
980
1030 1030 1032 1030
1110
1151
1201
1273
1353
1384 1382 1382 1384
1486
1620 1550 1545 1545
1640 1640 1645
1740
2830 2833 2832
3417 3476 3460 3468
3777 3776
96
rTrTrri
4000 3500 3000 2500 2000 1500 1000 cm
t
500
W ~ m b c r an-I
Figure 3B.4 Infrared spectra of SALSC complexes
97
A comparative study of the IR bands of the ligand and the complexes was
carried out. Some of the ligand absorptions are masked by the absorptions of the
zeolite. The ligand absorption at 1620cm-
1
due to uc=o is shifted to a lower region
(1550-1545cm-
1
) in the complexes. Azo-methine uC=N of the Schiff base, seen in
the region 1485cm-
1
[38] might have shifted to lower regions in the complexes,
though this is not clearly visible in the case of all the complexes, possibly because
it gets merged with other peaks. The band appearing at 1273cm-
1
is assigned to
the vibration of the phenolic (=0 group [39]. The shift of this band is not seen in
the spectrum of the complexes as it is masked by the broad band of the zeolite. The
broad band seen around 3250 _3500cm-
1
in the ligand can be due to the merging of
uO-H and uN-H frequencies. The uO-H and uN-H absorptions observed in the
spectra of the encapsulated SALSC complexes indicate the presence of the water
molecules of zeolite, and the uncoordinated NH2 group [40].
3B.4.7. EPR spectra
The EPR spectrum ofCuYSALSC is shown in Figure 3B.5 and the EPR
parameters are given in Table 3B.6.
Table 3B.6 EPRparameters ofCuYSALSC
EPR parameters Values
gll 2.34
All
179 x 10-4 cm-
1
gl/A
II
130cm
g.l
2.04
A.l
54.98 x 10-
4
cm-
1
~
0.8940
~
Ileff
1.85 BM
P 0.5589
98
R
5(,Q,C,! : ~ ,c111.
Figure 3B.S EPR spectrum of Cu YSALSC
The gll of CuYSALSC is 2.34 which is slightly higher than that expected
for square planar complexes. The gill I ratio is in the range reported for square
planar complexes. The EPR of neat CuCISALSC is reported to be of typical
distorted square planar geometry [35]. This indicates that there is a slight change
in the structure on encapsulation, which can be due to restriction imposed by the
zeolite cavity. The a
2
value of 0.894 suggests an ionic environment for CU(II) ions.
The magnetic moment of 1.8 BM agrees with that obtained
susceptibility measurements.
3B.4.8. TG analysis
in Figure 3B.6 and the data are presented in Table 3B. 7. Thermal studies of neat
SALSC complexes have been made by Petrovic et aL [41]. Well-defined TG
curves are obtained for neat complexes. Such types of curves are not obtained in
99
Table 3B.7 Thennal decomposition data
"" ...... '""" ...-.. .. .....-..---
---------
CoYSALSC NiYSALSC CuYSALSC
. ~ ~ . ~ n __ ......................................... _ ....................................... _ ....................... _ ........... __ ............................................. ___ .. _ .... _ ... _ ........... _ .....
Temp
Range
( QC)
30-100
100-250
250-500
500-690
..
% Temp
Mass Range
loss
CC)
5.24 40 -192
5.93 192 - 250
2.30 250- 505
0.33 505-700
I -c.nALSC I
%
Mass
loss
7.50
1.30
3.00
0.45
.00 \A
..
.. J
,. \
.,.
1.' ..
it
..
..
..
Temp
Range
( QC)
32-149
149-216
216-325
325-700
o lOCI 2'110:110.0 SO" MO ..... O IlOO
,_,.,......('1:)
"'.---------,
_c.YSoItoLSC
OD
..
..
..
o taO HO ~ o 0lil0 5:10 1(10 JOO ... to.
Tem,_ ... .,,. ('c)
Figure 3B.6 TG curves ofSALSC complexes
%
Mass
loss
9.40
2.40
0.87
1.13
.,s_se I
100
the case of encapsulated complexes. The stages are continuous rather than distinct.
This can be due to the very low concentration of the complex in the zeolite
support.
The weight loss due to the decomposition of ligand does not occur at a
distinct stage. This may be due to the very low concentration of the complex inside
the zeolite pore. However the mass loss in the temperature range 30-100 QC in all
the complexes can be due to loss of free water; that between 100 and 500 QC can
be due to the removal of coordinated water and the decomposition of the complex.
The mass loss between 500 and 670 QC can be due to the partial decomposition of
the complex and also due to the partial decomposition of zeolite.
101
References
1. M.A. Ali and S.E. Livingston, Coord. Chem. Rev. 13 (1974) 101.
2. D.H. Jones, R. Slack, S. Squires, and K.RH. Wooldridge, J. Med. Chem. 8
(1965) 676.
3. RW. Brockman, lR Thomson, MJ. Bell, and H.E. Skipper, Cancer Res. 16
(1956) 167.
4. Suhas A. Chavan, D. Srinivas, and Paul Ratnasamy, Chem. Commun. (2001)
1124.
5. Carmen Schuster, Eugen M611mann, Andras Tompos, and Wolf gang F.
H6lderich, Catal. Left. 74, 1-2 (2001) 69.
6. M. Kumar and S. Arabinda Kumar, Asian J. Chem. 6, 4 (1994) 782.
7. M. Kumar, Asian J. Chem. 6,3 (1994) 576.
8. C. Tollman and H. Herron, Symposium on Hydroc. Oxidation, 194th
National Meeting of American Chemical Society, New Orleans, LA, Aug. 30
- Sep. 4, 1987.
9. P.G. Menon, in Lectures in Catalysis, 4r
t
Ann. Meeting, Ind. Acad. Sci.,
S. Ramasheshan (Ed.) 1975.
10. Carol A Bessel and Debra R Rolison,J. Phys. Chem. BIOI (1997) 1148.
11. E. Paez-Mozo, N. Gabriunas, F. Luccaccioni, D. D. Acosta, P. Patrono, A
La Ginestra, P. Ruiz, and B. Delmon, J. Phys. Chem. 97 (1993) 12819.
12. H. Van Koningsveld, le. Jansen, and H. Van Bekkum, Zeolites 10 (1990)
235.
13. N.Herron,J. Coord. Chem.19(1988)25.
14. E.M. Flanigen, in Zeolite Chemistry and Catalysis, lA Rabo (Ed.) ACS
Monograph, American Chemical Society, Washington D.e., 1976.
15. P.A Jacobs, H.K. Beyer, and l Valyon, Zeolites 1 (1981) 161.
16. K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination
Compounds (5
th
edn.), John Wiley and Sons Inc., New York, 1997.
17. Bindu Jacob, Ph.D. Thesis, Cochin University of Science and Technology,
1998.
18. K. Klier, J. Am. Chem. Soc. 91 (1969) 5392.
19. U. Sakaguchi and AW. Addison, J. Chem. Soc. Dalton Trans. (1979) 600.
20. 1. Strutz, H. Diegruber, N.I. Jaeger, and R Mosler, Zeolites 3 (1983) 102.
21. K. Mizuno and lH. Lunsford,Inorg. Chem. 22 (1983) 3483.
22. T.A Egerton, A. Hagan, F.S. Stone, and le. Vickerman, J. Chem. Soc.
Faraday Trans. 168 (1972) 723.
23. 1.M. Thomas and e.RA Catlow, Progr. Inorg. Chem. 35 (1988) 1.
24. R. Raja and P. Ratnasamy, J. Catal. 170 (1997) 244.
25. Fethi Bedioui, Lionel Roue, Emmanuel Briot, Kenneth J. Balkus, Jr., and 1.
Felipe Diaz, New J. Chem. 20 (1996) 1235.
102
26. K.J. Balkus, Jr., and AG. Gabrielov, J. Inclusion Phenom. Mol. Recogn.
Chem. 21 (1995) 159.
27. S.P. Varkey and C.R Jacob, Ind. J. Chem. 37 A (1998) 407.
28. F.A Cotton and G. Wilkinson, Advanced Inorganic Chemistry, Wiley -
Interscience, New York, 1980.
29. A Earnshaw, Introduction to Magnetic Chemistry, Academic Press, London,
1968.
30. N.N. Greenwood and A Eamshaw, Chemistry of Elements (1 st edn.),
Pergamon Press, Oxford 1984. Rpt.1989.
31. AB.P. Lever, Inorganic Electronic Spectroscopy, EIsevier,
Amsterdam, 1968.
32. AB.P. Lever, G. London, and P.J. McCarthy, Can. J. Chem. 55 (1977) 3172.
33. B.Singh, B.P. Yadava, and RC. Agarwal, Ind. J. Chem. 23A (1984) 441.
34. H. Diegruber, P.J. Plath, G. Schulz-Ekloff, and M. Mohl, J. Mol. Catal. 24
(1984)115.
35. J Patole, S. Dutta, S. Padhye, and E. Sinn,Inorg. Chim. Acta 318 (2001)
207.
36. P.L. Maurya, B.v. Agarwala, and AK. Dey, Ind. J. Chem.19A (1980) 807.
37. R.N. Ray and B.K. Mohapatra, Ind. J. Chem.19A (1980) 590.
38. R.M. Silverstein and G.c. Bassler, Spectrometnc Identification of Organic
Compounds (4
th
edn.), John Wiley, New York, 1968.
39. Y.M. Leovac, L. Bjelica, and L. Jovanovic, Polyhedron 4,2 (1985) 233.
40. P.K. Singh, JK. Koacher, and JP. Tandon, J. Inorg. Nucl. Chem. 43 (1981)
1755.
41. AF. Petrovic, D.M. Petrovic, Y.M. Leovac, and M. Budimir, J. Thermal
Studies and Calonmetry 58 (1999) 589.
103
Chapter 4
A comparison of neat and encapsulated complexes of
copper 2-styrylbenzimidazole
Abstract
Neat and zeolite Y encapsulated copper 2-styrylbenzimidazole complexes were synthesized and a
comparative study of the two complexes was undertaken. Swface area, pore volume, and XRD
studies provide evidence for the encapsulation of the complexes without the destruction of the
zeolite crystallinity. The TG-DTG curves gave infOlmation regarding the thermal stability of the
complex and the probable composition of the expelled group. The formation of the complex was
confnmed by IR spectroscopy. Magnetic moment suggests square planar geometry for both the
neat and encapsulated complexes. Electronic spectra suggest the possibility of d-d transitions in
both the complexes. Square planar structure is confIrmed by the gill All ratio from EPR.
4.1. Introduction
Benzimidazole forms a significant part of the structure and function of a number of
biologically important molecules [1-4]. Derivatives of benzimidazole with
antifungal [5], anthelmintic [6], anti-lllV [7], antihistaminic [8-9], cardio tonic
[10], antihypertensive [11 ,12], neuroleptic [13] and such other pharmacological
effects are in clinical use. Several reports reveal that substitution at the 1, 2, and 5
positions of the benzimidazole ring. may influence their pharmacological effect
[14,15].
2-Substituted phenylbenzimidazole with antibacterial [16], antiviral [17],
antitumoral [18,19], anti-inflammatory [20] and other types of biological activities
have been reported. 2-Subsituted styrylbenzimidazole complexes of CU(II) were
prepared, in order to study the influence of steric effects on catalysis reaction. The
complexes may also be interesting with regard to their biological activities, even
though no attempt was made to study their biological properties.
Figure 4.1 Structure of2-styrylbenzimidazole
4.2. Experimental
Details regarding the synthesis and purification of the ligand
styrylbenzimidazole and the preparation of CU(II) exchanged zeolite Y are given in
Chapter 2.
105
4.2.1. Synthesis of neat Cu(SBhCh
CU(II) chloride dihydrate (10 mmol) was dissolved in methanol (25 ml) and
added to a solution of styrylbenzimidazole (20 mmol, 25ml) in methanol. This
mixture was refluxed for 3-4 hours. It was kept in the refrigerator overnight.
Crystals of styrylbenzimidazole complex that separated out were filtered, dried in
vacuo, and finally recrystallised from methanol to obtain pure styrylbenzimidazole
complex.
4.2.2. Synthesis of zeolite encapsulated 2-styrylbenzimidazole complex
The general procedure for encapsulating the complex in zeolite Y is
described in Chapter 2. Copper(II) exchanged zeolite, CuY (3g), was mixed with
excess of ligand (ligand to metal mole ratio > 2) and introduced into a glass
ampule, sealed, and heated at 150C for four hours. The resulting material was
purified by soxhlet extraction using methanol for 24 hours. Uncomplexed metal
ions present in the zeolite was removed by back exchanging with NaCI solution
(O.OIM in 250cc). This was made chloride free by washing with distilled water.
The zeolite encapsulated complex was stored in vacuum over anhydrous calcium
chloride after drying at 110C for 2 hours.
4.3. Characterization techniques
The various physico-chemical techniques to characterize the prepared
complexes are given in Chapter 2.
4.4. Results and Discussion
Zeolite Y encapsulated copper styrylbenzimldazole complex and neat
copper styrylbenzimidazole were synthesized. The neat complex is crystalline,
non-hygroscopic, green in colour, and is stable to aerial oxidation. The complex is
106
soluble in methanol, ethanol, acetone, and DMSO, and is insoluble in solvents like
chloroform, benzene, hexane, and dichloromethane. The molar conductance of the
complex measured in 10-
3
M methanol solution is in the range of 1.1 ohm-
l
cm
2
morl indicating its non-electrolytic nature [21].
4.4.1. Chemical analysis
The results of chemical analysis of styrylbenzimidazole complexes are given
in Table 4.1. The data obtained in CRN analysis, metal analysis, and chlorine
analysis are in good agreement with the calculated value of carbon, hydrogen,
nitrogen, copper, and chlorine for the empirical formula, CuClz(SBh. Chemical
analysis of CuYSB shows that the Si/AI ratio is 2.4, which is the same as the Si/
AI ratio ofNa Y and CuY, indicating retention of the zeolite framework. The metal
content in Cu Y is 1.12% and that in Cu YSB is 0.45%. About 40% of the metal
initially present in the metal exchanged zeolite was found to remain in the zeolite
after complexation and the rest was removed during the final ion exchange with
NaCI solution.
Table 4.1 Analytical data of styrylbenzimidazole complexes
Sample % % % % % % % %
Si AI Na C H N Cl Cu
NaY 20.60 8.50 7.50
CuY 20.48 8.48 6.60 1.12
CuYSB 20.48 8.40 6.20 0.30 2.40 0.04 0.45
Cu(SBhClz*
61.96 4.l8 9.33 11.80 10.50
Cu(SBhCh 62.56 4.20 9.70 12.30 11.03
------
Cu(SBhClz* : Found Cu(SBhCh**: Calculated
107
The charge neutralization of the complex in the zeolite cage is probably
achieved by co-ordination of the metal sites to the oxide ions of the lattice. The
ligand to metal mole ratio in the neat complex is 2:1. But the mole ratio is less
than 2 in the case of the encapsulated complex. The lower mole ratio of the zeolite
complex may be because of the presence of un-exchanged metal ions in the lattice,
which have not been removed, in the final ion exchange with sodium chloride.
The encapsulated complexes might have kept these metal ions trapped, thus
preventing ion exchanging with Na + ions.
4.4.2. X-ray diffraction patterns
Figure 4.2 shows the XRD patterns of CuY and CuYSB. The XRD patterns
of the encapsulated complex exactly match those of the copper exchanged zeolite.
This shows that the crystalline structure is retained even after encapsulating the
complex within the lattice.
4.4.3. Scanning electron microscopy
The scanning electron micrographs (Figure 4.3) before soxhlet extraction
and after soxhlet extraction indicate the removal of surface particles.
4.4.4. Surface area and pore volume
Surface area and pore volume data are given in Table 4.2. The surface area
and pore volume of metal exchanged zeolite is higher compared to those of the
encapsulated complex. The lowering of the surface area and pore volume of the
encapsulated complex can be due to the filling of the zeolite pore with metal
complex. This is again strong evidence for the encapsulation of the complex in
zeolite pore [22].
108
Figure 4.2 XRD patterns
i) CuYSB and ii) CuY
(ii)
IC
"';L:l'rt7:t.;;
109
(i)
(ii)
Figure 4.3 SEM of CuYSB
i) Before soxhlet extraction
ii) After soxhlet extraction
110
Table 4.2 Surface area and pore volume data
. - - - - - - ~ . . . . . . . - - ~ - - - - - - - - - - - - - -
Surface area (m /g) Pore volume (mllg)
CuY CuYSB % loss CuY CuYSB % loss
520 150 71.2 0.293 0.069 76.3
4.4.5. Magnetic moment
The magnetic moment for CuCh(SBh and CuYSB at room temperature are
l.80 and 1.81 BM. respectively. This indicates that the complexes are monomeric
in nature and may have square planar or distorted square planar structure [23].
Further the values suggest absence of any metal-metal interaction in the complexes
[24].
Table 4.3 Magnetic moment
Sample
CuCh(SBh
CuYSB
4.4.6. Diffuse reflectance spectra
Magnetic moment
(BM)
1.80
1.81
The electronic spectral data and the tentative assignment are given in Table
4.4, and the spectra are given in Figure 4.4. The electronic spectra of the neat and
encapsulated complexes show a band in the region 12500cm-
1
which could be
111
assigned to d-d transitions. The band around 29400cm-
1
in the spectra of the neat
and encapsulated complexes can be attributed to a charge - transfer transition
[25,26].
Table 4.4 Electronic spectral data of copper complexes of styrylbenzimidazole
I
Xl
Sample Abs. max. Assignment
( -1\
_________________ IE __ L ____ _
12500
CuYSB 29400
12500
Cu(SBhCh 29400
~
-..'
,
~
!ID !ID
VIIiM!Iergth (mI
d-d transition
Charge transfer transition
d-d transition
Charge transfer transition
Q3)
Q25
~
Q22
Figure 4.4 Electronic spectra of SB complexes
112
Square planar geometry can be assigned to both neat and encapsulated complexes.
In the case of the neat complex, two of the four ligands are chloride ions, whereas
in the case of the zeolite encapsulated complex they can be oxide ions of the
zeolite framework or molecules of water.
4.4.7. FTIR spectra
Figure 4.5 shows the IR spectra of the styrylbenzimidazole ligand, and its
neat and encapsulated complexes. The spectral data of the styrylbenzimidazole
ligand, the neat complex, and of the encapsulated complex are presented in Table
4.5. From the table, the ligand bands in the encapsulated complexes which are not
masked by zeolite bands can be readily identified. The band of medium intensity at
1516cm-
1
in the spectrum of the free ligand SB may be attributed to the C-N
stretching vibration. In the case of the neat and encapsulated complexes this band
shifts to 1500 cm-I. The shifting of this band to lower frequency is due to the
coordination of nitrogen of the SB to the metal atom. The broad band at - 3500cm-
I can be attributed to the stretching vibration of water.
4.4.8. EPR spectra
EPR spectra of the neat copper styrylbenzimidazole (glass spectra) and the
encapsulated complex recorded at liquid nitrogen temperature are shown in Figure
4.6. Table 4.6 gives the EPR parameters, unpaired electron density, and magnetic
moment, calculated from EPR data. The EPR ofCuClz(SB)z in DMSO shows a set
of four copper hyperfine lines on the gll position in the low field range, while such
hyperfine lines are not so conspicuous in the KL position in the high field region.
No super-hyperfine splitting from the ligand is observed either. The gll and
values of CuCh(SB)2 are 2.32 and 2.05, and those of the zeolite encapsulated
complexes are found to be 2.34 and 2.05 respectively.
113
Table 4.5 IR spectra of the styrylbenzimidazole ligand and its complexes
........................................................................ "" ......... "" ............................................................................................... """ ............................................................................................. u. ....
SB CU(SB)2Ch CuYSB
cm
-\
cm
-\
cm
-\
433 430 450
495 480
699 700
750 760 780
845 820
903 890
965 960
1020 1040
1151 1140
1182 1200
1227 1250
1340
1381
1423 1440
1485
1516 1500 1500
1558 1550
1588
1645 1640 1640
1946 1900
2608 2600
2731 2720
2808
2849 2880
2971
3054 3040
3180
3400
~ ~ ...--...---.
114
(ii)
n l ~
QJ
f
()

...
I
I
....
S
IZl
s::
~
~
(i)
/
~
r
i
I
4000 1500
o
Wavenumber cm-
1
Figure 4.5 IR spectra of the styrylbenzimidazole ligand and its complexes
(i) styrylbenzimidazole, (ii) Cu(SB) 2Ch, (iii) CuYSB
115
The gllA" value ofCuCh(SBh is 119.6 cm while that ofCuYSB is 122.S
cm, suggesting a square planar geometry for the neat and the encapsulated
complexes. The a? values of 0.72 and 0.8 for the neat and encapsulated complexes
suggest that the metal ligand bond is intermediate between ionic and covalent. The
trend g" > 8.L > 2.0023 observed for these complexes indicate square planar
structure for these complexes. The magnetic moment, !leff for CuCh(SB)2 and
CuYSB are1.84 and 1.86 BM, which is in agreement with the value obtained from
room temperature magnetic susceptibility measurement by Gouy method [27].
The well resolved hyperfine splitting similar to that for the neat complex in dilute
solution indicates that the complex species in the zeolite cavity are monomeric in
nature.
Table 4.6 EPR parameters
--_.,-----
EPR parameters Cu(SBhCh CuYSB
gll
2.32 2.34
A"
194x 10-4
cm
-! 191 x 10-4
cm
-!
g11/A
11
119.6cm 122.Scm
8.L
2.0S 2.0S
A.l
36 x 10-4 cm-! 34 x 10-4 cm-!
a?
0.72 0.8
Ileff
1.84 1.86
116
(iii)
(ii)
(i)
3000
Figure 4. 6 EPR parameters of SB complexes of copper
i) CuYSB at LNT ii) Cu(SB)zClz at LNT iii) CuYSB at RT
117
4.4.9. TG analysis
The TGIDTG curves of CU(SB)2Ch were recorded in an atmosphere of
nitrogen from room temperature to 1200 C at a heating rate of 10 o/min and the
TG curve of the encapsulated complex was recorded in air from room temperature
to 800C at a heating rate of 10C/min. Figure 4.6 (i) shows the TGIDTG curves
of Cu(SBhCh and Figure 4.6 (ii) shows the TG curve of CuYSB. Thermo
analytical data for the complexes are presented in Table 4.7. The percentage mass
loss and the probable composition of the expelled groups are also given in this
table.
Table 4.7 Thermal decomposition data
Peak Temp. % Mass loss Probable
temp. range from composition of
Sample DTG TG Stage TG expelled group
(0C)
(C) Found
(Calc.)
396 202-551 I 40.26 SB
(38.42)
Cu(SBhCh
737.4 551-975 IT 48.26 SB + 2Cl
(50.73)
30 - 100 5.36 Water
100 - 534 12.87 Water and
CuYSB ligand
534 -799 0.5391 Ligand and
zeolite
_ ................ -....................... _-----_ .. _------_ ...""""' ...........""'"
--,---,-
118
co
2.
130
-, ... -.... - ..... """ ...
".
, .
\ :
1
I,
'.'
1"
330
- -
(ii)
,,,..._ ... , .... - ~ " ~ ' , - ... -..... ,
.
I
,
,
,
(i)
Figure 4.7 (i) TGIDTG curves of Cu(SB hCh and (ii) TG curve of
CuYSB.
In the case of Cu(SB)2Ch the mass loss in the first stage corresponds to the
expulsion of one styrylbenzimidazole. The IR spectrum of the residue after this
stage showed the presence of styrylbenzimidazole indicating that only one
styrylbenzimidazole is removed. The elemental analysis after the first stage
indicated the approximate composition of the intermediate complex to be
Cu(SB)Cb. The second stage of decomposition corresponds to the removal of one
119
styrylbenzimidazole and two chlorine atoms. The residue after the final stage of
decomposition was found to be Cu
2
0 (ultimately oxidized to CuO).
From the TG curve of the encapsulated complex we get only a qualitative
idea. The curve shows that most of the intrazeolite free water molecules are
released in the temperature range 30-100 0c. The mass loss due to the expulsion of
the ligand is not clearly visible as in the case of the neat complex. The mass loss
in the temperature range 100-534 C can be due to the removal of ligands and
coordinated water if any. The absence of well-defined patterns as in the case of
neat complexes can be due to the very low concentration of the metal complexes
inside the zeolite pore. The mass loss corresponding to the temperature range
between 534 C and 800 C can be due to the loss ofligand and the decomposition
ofthe zeolite itself. The IR spectrum of the residue at the end ofthis stage
indicates that the encapsulated complexes have decomposed completely.
120
References
1. M.N. Hughes, The Inorganic Chemistry o/Biological Processes (2
nd
edn.),
Wiley, London, 1981.
2. R W. Hay, Bio-inorganic Chemistry, Ellis HOlwood Limited, England,1984.
3. H.c. Freeman, in Inorganic Biochemistry, G.L. Eichhom (Ed.), Elsevier,
Armsterdam, 1973.
4. Cannon KUS, Turk. J. Chem. 26 (2002) 559.
5. D. Berg, K.H. Buchel, M. Plempel, and A Zywietz, Mykosen 29 (1986) 221.
6. AG. Salmot, AC. Cremleux, J.M. Hay, A Meulemans, M.D Glovanangeli,
B. Delaitre, and F.P. Coulaud, Lancet 17 (1983) 652.
7. A Chimirri, S. Grasso, AM. Monforte, P. Monforte, and M. Zappala,
Il Farmaco 46 (1991) 925.
8. C.1.E. Niemegeers, F. Awouters, and P.AJ. Jannsen, Agents and Actions 18
(1986) 141.
9. R lemura, M. Hori, and H. Ohtaka, Chem. Pharm. Bull. 37 (1989) 962.
10. K. Ishihara, T. Ichikawa, Y. Komuro, S. Ohara, and K. Hotta, Arzneim.-
Forsch./Dntg Res. 44 (1994 ) 827.
11. D.Y. Graham, A Mccullough, M. Sklar, J. Sontag, W.M. Roufail, R Stone,
RH. Bishop, N. Gitlin, AJ. Cagliola, RS. Berrnan, and T. Humphries,
Digetive Diseases and Sciences 35 (1990) 66.
12. G. Piazzesi, L. Morano, and J. C. Riiegg, Arzneim.-Forsch.,IDrug Res. 37
(1987) 1 NI.
13. I. Wiedemann, H. Peil, H. Justus, S. Adamus, V. Brand, and H. Lohmann,
Arzneim.-Forsch.,IDntg Res. 35 (1985) 964.
14. H. Goker and C. Pharm. (Weinheim) 328 (1995) 425.
15. L. Garuti, M. Roberti, and G. Gentilomi, Il Farmaco 55 (2000) 35.
16. RA Cobum, M.T. Clark, RT. Evans, and RJ. Genco, J. Med. Chem.30
(1987) 205.
17. T. Roth, M.L. Momingstar, P.L. Boyer, S.H. Hughes, RW. Buckheit, and
C.1. Michejda, J. Med. Chem. 40 (1997) 4199 .
18. AY. Chen, C. Yu, A Bodley, L.F. Peng, and L.F. Liu, Cancer Research 53
(1993) 1332.
19. W.A Denny, C.W. Rewcastle, and B.c. Baguley, J. Med. Chem. 33 (1990)
814.
20. D. Evans, T.A Hicks, W.RN. Williamsom, W.Dawson, S.C.R Meacock,
and E.A Kitchen, Eur. J. Med. Chem. 31 (1996) 635.
21. WJ. Geary, Coord. Chem. Rev. 7 (1971) 81.
22. E. Paez-Mozo, N. Gabriunas, F. Lucaccioni, D.D. Acosta, P. Patrono,
AL.Ginestra, P. Ruiz, and B. Delmon,J. Phys. Chem. 97 (1993) 12819.
121
23. B.V. Romanovsky and A.G. Gabrielov, J. Mol. Catal. 74 (1992) 293.
24. YM. Issa and W.H. Hegazy, Synth. React. In 0 rg. Met.- Org. Chem. 30, 9
(2000) 1731.
25. K. Krishnankutty and 1. Michael, J. Indian Chem. Soc. 72 (1995) 461.
26. P.G. Edward, 1.S. Fleming, and S.S. Liyanage, J. Chem. Soc. Dalton Trans. 2
(1997) 193.
27. N.N. Greenwood and A. Earnshaw, Chemistry of Elements (18t edn.)
Pergamon Press, Oxford, 1984. Rpt.1989.
122
Chapter 5
Zeolite encapsulated dithiocarbamates of Fe, Co, Ni and eu
Abstract
Fe(ill) dithiocarbamates of MDTC, IPDTC, and EPDTC and Co(Ill), Ni(II) and Cu(II)
carbodithioates of MDTC were encapsulated in the super cages of zeolite Y. Si/AI ratio, XRD
patterns, surface area, and pore volume data suggest the encapsulation of the complex without
affecting the crystalline structure of the zeolite. SEM shows a clean surface. Magnetic moment and
electronic spectra indicate octahedral structure for the Fe(III) complexes and Ni(II) and Co(II)
complexes. The Cu(lI) complex has square planar geometry. This is further confIrmed by the g"/A,,
ratio. Different stages are not distinct in the TG curves and they are continuous.
5. Introduction
The coordination chemistry of dithiocarbamates has been a subject of great
research interest for a number of years. Dithiocarbamates form a wide range of
complexes with transition metals. The most investigated group of dithiocarbamates
is that derived from secondary amines because of their stability and redox
properties. Another significant fact is that dithiocarbamates can act as both
monodentate and bidentate ligands.
(i)
(ii)
(iii)
Figure 5.1 Structure of the dithiocarbamate ligands
i) morphine-N-carbodithioate (MDTC), ii) di-iso-propyldithiocarbamate (lPDTC),
and iii) N-ethyl-N-phenyldithiocarbamate (EPDTC)
124
The dithiocarbamate anion has a 4-electron-three-centre rr system. The
bonding of the ligand to transition metals is through the two S atoms, and both the
M-S distances are equal. In a planar MS2CN system, an extensive rr-delocalization
exists with a high contribution to the resonance structures, and with relatively high
electron density on the metal [1]. The peculiar properties of dithiocarbamates are a
consequence of their high covalency and special type of electronic ground state
de localization. Schrauzer attributes this delocalization to the interaction of the
lowest occupied rr-MOs of the ligands, with the metal p-orbitals of appropriate
symmetry, producing a low-lying molecular orbital delocalized over the whole
molecule or complex [2]. Researchers have explored the role of dithio ligands in
the design of several electrically conducting molecular solids. The current interest
is due to their structure [3], electrochemical properties [4], and photo-redox
property [5]. Further, metal dithiocarbamato complexes find application as
accelerators in the vulcanization of rubber. It was therefore considered worthwhile
to encapsulate the metal complexes of Fe(III), Co(III), Ni(II) and CU(II)
dithiocarbamates inside zeolite Y by the flexible ligand method to study the effect
of encapsulation on the geometries, and on the magnetic, spectral, and catalytic
properties. This chapter is divided into 2 sections. Section A deals with the
synthesis and characterization of Fe(III) dithiocarbamates; and Section B deals
with the synthesis and characterization of Co(III) , Ni(II). and CU(II)
morpholinecarbodithioates.
Section A
Zeolite-Y encapsulated Fe (Ill) dithiocarbamate and carbodithioate
complexes
SA.1. Introduction
Iron dithiocarbamates and carbodithioate complexes in the oxidation states
II, Ill, and IV have been studied extensively. It has been reported that Fe(R2DTC)2
125
is unstable and quickly gets oxidized in air to Fe(R2DTC)3. The redox properties of
tris chelated Fe(III) have been investigated in detail [8-11]. Iron dithiocarbamates
have widespread use as spin trap agents for nitric oxide detection. However these
complexes have not been encapsulated in zeolite Y cages and studied.
Encapsulation is found to impart improved properties to many systems, which
have been explored for various applications. The results of the studies on zeolite Y
encapsulated Fe(III) dithiocarbamates are presented in this section.
SA.2. Experimental
The details regarding the synthesis and characterization of the ligands
morpholine-N-carbodithioate (MDTC), N-ethyl-N-phenyldithiocarbamate
(EPDTC) and di-iso-propyldithiocarbamate (IPDTC) are presented in Chapter 2.
SA.2.t. Synthesis of zeolite encapsulated iron carbodithioate and
dithiocarbamates
Iron(III) exchanged zeolites (3 g) is mixed with excess dithiocarbamate
ligands in 50 ml of water (the concentration was adjusted to have a ligand to metal
ratio greater than 3) and it was refluxed for 5 hours. Similar procedure has been
reported for the synthesis of Fe(I!) complex in HEU zeolite with N,N,-
diethyldithiocarbamate [13]. The complex was washed several times with water to
remove any uncomplexed ligand, and soxhlet extracted using methanol for 24
hours to remove surface species. These were finally sodium exchanged using
NaCl solution (O.OlM. 250 ml), washed free of chloride, dried at 110 QC, and
stored in vacuum over anhydrous CaCh.
SA.3. Characterization techniques
Details of the various physicochemical techniques for the characterization
of the complexes are given in Chapter 2.
126
SA.4. Results and Discussion
SA.4.1. Chemical analysis
The analytical data of encapsulated Fe(JIT) complexes are given in Table
SAl. The Si/AI ratio of these complexes was found to be 2.4 as in the case of the
metal exchanged zeolites.
Table SA.1 Analytical data of the encapsulated Fe(III)
carbodithioate and dithiocarbamates
Sample % Si % AI % Na % M %C
............................................................................................................................................................
FeYMDTC 18.48 7.83 5.40 0.53 1.56
FeYEPDTC
FeYIPDTC
18.40 7.78
18.70 7.91
5.55
5.80
0.51 2.90
0.50 2.25
The metal to ligand ratio in neat iron dithiocarbamates is 1:3 [14,15]. The
analytical data for the complexes suggest that the percentage of metal is in the
range as that required for 1:3 complex (with slight excess of metals).
SA.4.2. X-ray diffraction patterns
XRD provides information on crystallinity [16,17] as well as change in unit
cell parameters that might result from encapsulation. XRD patterns ofFeYMDTC,
FeYEPDTC, and FeYIPDTC are given in Figure 5A.1.
127
5 1.0 15 20 25 30 35 4Q
Figure SA.! XRD patterns of encapsulated iron dithiocarbamate complexes
i) FeYMDTC, ii) FeYlPDTC, iii) FeYEPDTC
The XRD pattern obtained for the ion exchanged and encapsulated complexes
were similar in nature. This indicates that there is no change in the overall
crystallinity of zeolite Y lattice on encapsulation.
5A.4.3. Scanning electron micrographs
The scanning electron micrographs ofFeYMDTC before soxhlet extraction
and after soxhlet extraction are presented in Figure SA2. The picture shows a
clean surface compared to that before soxhlet extraction.
128
(i)
Figure SA.2 SEM
i) Before soxhlet extraction
ii) After soxhlet extraction
129
SA.4.4. Surface area and pore volume
Surface area and pore volume data are given in Table 5A.2. The surface
area and pore volume of metal exchanged zeolites were found to be lowered on
encapsulation. The lowering of surface area and pore volume confirms the
encapsulation of the complex in the zeolite pores.
Table SA.2 Surface area and pore volume data
' ' ' ' ' ' ' ' ' - ' - ' ' ' ' - - - - ' ' ' ' ' ' ' ' - - ' ' - - - - - - - - - ' ' ' - - - - ' - - - - - - ' - - ~ - - ' - ' - - , - , - - - - - - - - - - - - , - , - - - - - - - , - , - - - - , - - - - - - - - , , - - - - -
Surface area (mo/g) Pore volume (mIlg)
. ~ .. .................................................................................................................................. _ .......... _ ...... _ .... _ ............................................................................. _"" .......................... .
Sample FeY FeYL % loss FeY FeYL % loss
FeYMDTC
FeYIPDTC
FeYEPDTC
530
530
530
SA.4.S. Magnetic moment
257
236
230
51.50
55.45
56.67
0.2982
0.2982
0.2982
0.1390 53.39
0.1359 54.43
0.1300 56.41
The magnetic moments of Fe(III) dithiocarbamates at room temperature are
close to those observed for high spin Fe(III) octahedral complexes. The results
obtained from magnetic moment measurements are presented in Table 5A.3.
Table SA.3 Magnetic moment
Sample
FeYMDTC
FeYEPDTC
FeYIPDTC
Magnetic moment
(BM)
5.87
5.81
5.79
130
SA.4.6. Diffuse reflectance spectra
The reflectance spectral data and tentative assignments are given in Table
5A.4, and the reflectance spectra in Figure 5A.3. Similar spectra are seen in the
case of all the encapsulated Fe(III) complexes. The band around 14080-14100 cm-I
seen in all the complexes can be the result of the following three transitions, 4T Ig
+- 6
AIg
, 4
T2g
+- 6
AIg
" 4Eg +-- 6
AIg
[18,19] in octahedral symmetry. The bands at
20830 and 21000cm-
1
for all the complexes could be considered d-d bands [18].
Charge transfer transitions are seen in the high-energy region around 32790-
33330cm-
l
. The room temperature magnetic susceptibility measurements are fully
consistent with the high spin octahedral structure of Fe(III} complexes. The results
indicate that the structure of the neat complexes is retained in the zeolite cages.
Neat Fe(III) dithiocarbamates are reported to exist in octahedral geometry [14].
Table SA.4 Electronic spectral data
Sample Abs. max. Assignment
(cm-I)
"i4i ..
FeYMDTC 4
T2g
+-- 6
AIg
4E g+-- 6
AIg
20900 d-d transition
32800 Charge transfer transition
14080
4 6
T
Ig
+- A
Ig
FeYEPDTC
4T 6
2g+-- AIg
4 6
Eg+- A
Ig
20830 d-d transition
33330 Charge transfer transition
14080
4 6
T
Ig
+-- A
Ig
FeYIPDTC
4 6
T
2g
+-- A
Ig
4Eg+--6AIg
21000 d-d transition
33330 Charge transfer transition
I FeYIPDTC I
04

Wave length (nm)
I FeYMOTC I
D.4
0.2 .......
WaveJength (om)
..
l:!
i
..
..Q

I FeYEPDTC I

Wailelength (nm)
Figure SA.3 Electronic spectra
of encapsulated iron
dithiocarbamate complexes
132
SA.4.7. FTm spectra
Infrared spectra of morpholinecarbodithioate, N-ethyl-N-
phenyldithiocarbamate, di-iso-propyldithiocarbamate and the encapsulated
complexes are shown in Figures 5A4, 5A5, 5A6. IR. spectral data of J.\.1DTC,
EPDTC, IPDTC, and the Fe(III) encapsulated complexes are presented in Table
SA5. vN=CS
2
absorption occurs at 1440cm-
1
in the spectrum ofIPDTC, 1450cm-
1
in the spectrum of EPDTC, and 1460cm-
1
in the spectrum of J.\.1DTC. Similar
observations have been made by earlier workers [19]. This is shifted to higher
frequency regions on complexation. The bands are seen at 1480cm-
1
in the case of
FeYIPDTC, 1490cm-
1
in the case of FeYEPDTC, and 1485cm-
1
in the case of
Fe YMDTC. The positive shift in vN=CS
2
vibration may be indicative of the
interaction of the sulphur atoms with metal on coordination [20]. The absorption in
the region 941-1140cm-
1
in the spectra of the ligands may be attributed to
symmetric and asymmetric C-S stretching vibrations [21]. These absorptions shift
to lower frequency on complexation. Similar lowering is expected in the case of
the encapsulated complexes also. But the new bands are not observed in the
spectra of zeolite complexes as they might have been masked by the broad zeolite
absorption at around 1000cm-
1
. Absorption bands due to the O-H stretching
vibrations of the zeolites lie in the range of 3440-3470cm-
1
and are seen in all the
complexes [19]. Most of the absorptions of the ligands are masked by the zeolite
peaks.
133
Table SA.S IR spectral data of iron complexes
................................................... "" .............................................................. "' ..................... u. ............................................................................ _ .................................... u. ........ "" ............................................................................... --. ...... -.-_uo. ......................................
MDTC FeYMDTC IPDTC FeYIPDTC EPDTC FeYEPDTC
cm-! cm-! cm-! cm-! cm-!
cm
-!
.................................................................................................................................................................................................................
481
540 558
639 618 610 618 620 618
692
724 719 726
780 757
888 846 816
903
983 941 992
1033 1035 1035 1058 1033
1110 1143 1122
1213 1193 1235
1260 1301 1321 1270
1301
1363 1352 1364 1352 1392 1352
1416
1460 1485 1440 1480 1450 1490
1615 1609 1598
1635 1630 1642
2115 2070 2067
2373
2851
2908 2968 2969
3379 3420 3378
3453 3443 3468
134
(i)
I
;
1:
I
"
! i
'"



4000 3500 3000 2500 2000 1500 1000
0"'"'
SOD
Figures 5A.4 IR. spectra
i) lVIDTC and ii) FeYNIDTC
135
1000 3500 3000 Z5DO ZlJOO 1500 1000
Figures SA.S IR spectra
i) IPDTC and ii) FeYIPDTC
136
4000 3500 3000 2500 2000 1500 1000 ell< 500
Figures SA.6 IR spectra
i) EPDTC and ii) FeYEPDTC
137
SA.4.8. TG analysis
The TG curve of FeYIPDTC is shown in Figure SA7. The different stages
III the decomposition of the complex and the zeolite cannot be clearly
distinguished here.
Table SA.6 Thermal decomposition data
FeYIPDTC
Temp range (0C)
30 -101
101 - 401
401 - 799
% Mass loss
4.90
9.81
0.43
Free water associated with the zeolite is lost in the temperature range between
30C and 100C. The mass loss in the temperature range 101 - 401C can be
due to the loss of coordinated water and partial decomposition of the ligand. A
mass loss of 0.43% occurs in the range 401 - 799 C, which represents the partial
decomposition of the complex as well as the zeolite matrix. The complete
decomposition of the complex at 800C was confirmed by analyzing the residue
obtained, using infrared spectroscopy.
-4.906
14.52
..... f-.... ----.. ----- .. A .. ,._ .... - ....... - __ -., ... -_ ... _ ..... - .. - . -_ .. _ ...... _ ..... ,
i 1
( .9.111 I
_ .. _ .. _ .. _ .. --_-.. .. .. _ .. _1 I
...... ...... w ........ ... ..
-o.433Sf
-..... .. -
.
,8 3);)
r't';
Figure SA..7 TG curve ofFeYIPDTC
138
Section B
Zeolite encapsulated CoOn), Ni(n).,and Cu(n)
morpholine-N-carbodithioate complexes
5B.l Introduction
Co(III), Ni(II) and CU(II) morpholine-N-carbodithioate complexes have
been studied. However, the effects of encapsulating these complexes inside zeolite
Y cages have not yet been studied. In this section our studies on the synthesis and
characterization of the encapsulated morpholinecarbodithioate complexes are
presented.
5B.2. Experimental
The details regarding the synthesis and characterization of the ligand
morpholine-N-carbodithioate (MDTC), and the ion-exchanged zeolite are
presented in Chapter 2.
5B.2.1. Synthesis of zeolite encapsulated Co(ID), Ni(II) and CU(II)
morpholine-N-carbodithioate complexes
The general method for synthesizing zeolite encapsulated complexes is
given in Chapter 2. A mixture of metal exchanged zeolite and morpholine-N-
carbodithioate in the ligand to metal mole ratio greater than 3 was refluxed over a
water bath for 5 hours. The resulting complex was washed with water and soxhlet
extracted using methanol. This procedure was adopted for the synthesis of the
encapsulated Co(II), Ni(II), and CU(II) complexes. But the encapsulated cobalt
complex, which was green in colour, leached out during soxhlet extraction
139
(became colorless after soxhlet extraction). So a different procedure was used for
the encapsulation of Co YMDTC. Encapsulated CO (II) complex was oxidized to
Co(III) by passing air through an aerator while heating the solution over a water
bath at 90C. The process was continued until the colour of the zeolite complex
changed completely to blue. Co(III)YMDTC forms tris complexes with MDTC
and will have size large enough to be trapped in the super cage. This was washed
with water and soxhlet extracted using chloroform.
The uncomplexed metal remaining in the zeolite in each of the three cases
was removed by back-exchange of the zeolite with O.OIM NaCI solution for 24
hours. It was then washed with water to remove chloride ions, dried at 100C for 2
hours, and stored in vacuum over anhydrous calcium chloride.
5B.3. Characterization techniques
The various analytical techniques used for the characterization of the prepared
complex are given in Chapter 2.
5B.4. Results and Discussion
5B.4.1. Chemical analysis
Zeolite Y encapsulated MDTC complexes of Co(III), Ni(II) and CU(Il) were
synthesized. The analytical data of encapsulated MDTC are presented in Table
SB.I. The cobalt (II) complexes could not be encapsulated in the zeolite cages.
The complexes were found to leach out of the cage, which might be due to smaller
size of the his complex and also might be due to the labile nature of the complex.
The data reveal that the Si/AI ratio of the metal exchanged zeolites (2.4) remains
unchanged in zeolite complexes indicating the capability of the zeolite structure to
withstand encapsulation conditions. Analytical data suggest that almost all the
metal ions get complexed.
140
Table SB.l Analytical data
CoYMDTC 18.46
NiYMDTC 18.52
CuYMDTC 18.48
SB.4.2. X-ray diffraction patterns
7.80
7.70
7.56
5.32
5.60
5.75
0.25
0.50
1.01
0.76
1.50
1.90
The XRD patterns of CoYMDTC, NiYMDTC, and CuYMDTC are given
in Figure5B.I. The encapsulated materials exhibited well-resolved XRD patterns
identical to that of metal exchanged zeolites. The incorporation of the complex did
not alter the XRD patterns, indicating that no structural change or decrease in
crystallinity has taken place.
SB.4.3 Scanning electron micrographs
The scanning electron micrographs ofFeYMDTC before soxhlet extraction
and after soxhlet extraction were taken. The micrograph taken after soxhlet
extraction showed that most of the particles were removed by this process. This is
shown in Figure 5B.2.
141
~
'<iii)
(ii)
(i)
5 10 2.0 25 3.0 35 40
Figure 5B.l XRD pattern of i) CoYMDTC, ii) NiYMDTC and iii) CuYMDTC
142
(i)
(ii)
Figure 5B.2 SEM
i) Before soxhlet extraction
ii) After soxhlet extraction
t43
SB.4.4. Surface area and pore volume data
Surface area and pore volume data are given in Table 5B.2. The lowering
of surface area and pore volume on encapsulation is indicated in the table. Surface
area and pore volume analyses provide strong evidence for encapsulation.
Table 5B.2 Surface area and pore volume data
' - - - ' ~ ' ' ' ' - ' ' ' ' - - ' ' ' ' - - - ' ' ' - ' ' - - ' ' ' ' ' - - ' ' ' ' ' ' : l - - - - ' ' - - - - - - - ' ' - - ' - ' ' ' - - ' ' - - ' ' ' ' ' ' ' ' - ' ' - - - - ' ' , , , , , , , , , , , , , , , , , ,
Surface area (m /g) Pore volume (mUg)
-........ -.... _ ...... _--_ ...... _ .._--_ .._----------.._----_ .... _ ...... _----
Sample
MY MYMDTC % loss MY MYMDTC % loss
CoY 532 250 53.0 0.2986 0.1385 53.6
NiY 528 261 50.6 0.2974 0.1384 53.3
CuY 530 244 53.9 0.2983 0.1342 52.3
SB.4.S. Magnetic moment
The magnetic moment calculated from magnetic susceptibility data are
summarized in Table 5B.3. Co(UI)Y1'vIDTC is found to be diamagnetic. The
magnetic moment 3.20 BM observed for NiY1'vIDTC indicates an octahedral
geometry for the Ni(n) ion, even though a neat complex of nickel dithiocarbamate
has a square planar geometry [22]. Cu Y1'vIDTC exhibits a magnetic moment of
1.82 BM, which suggests a square planar geometry for the complex.
Table 5B.3 Magnetic moment
Sample Magnetic moment
CoYMDTC
NiYMDTC
CuYMDTC
(BM)
3.20
1.82
144
5B.4.6. Diffuse reflectance spectra
The electronic spectral data of encapsulated MDTC complexes are given in
Table 5B.4 and the spectra are given in Figure 5B.3. The spectrum ofCoYMDTC
shows bands at 16000cm-
1
and 24390cm-
l
. These absorptions agree with those
reported for octahedral complexes ofCo(III) [18]. These bands could be attributed
to IT g +- I Ag and IT 2g +- I Ag transitions in octahedral Co(IIJ) complexes. The
transition at 30120cm-
1
is a charge transfer transition.
Table 5B.4 Electronic spectral data
. ~ ~ ~
Sample Abs. max. Assignment
CoYMDTC
NiYMDTC
CuYMDTC
(cm-I)
16000
24390
30120
15000
25000
31060
14900
17000
28570
I I
T
lg
- A
Ig
IT2g_IAlg
Charge transfer transition
3
T
I g(F) _ 3 A2g(F)
3
TIg
(P) _ 3
A2
g(F)
Charge transfer transition
B
2
g -BIg
Alg-Blg
Charge transfer transition
145
CD
i
o
..
,g
0(
ICoYMDTCI
0.5
400 600 BIlO
Wavelength (nm)
0.25
ICuYMDTCI
0.20
0.15
o 1 -"'------:"800
Wavelength (nm)
0.21
I NiYMDTcl
0.20
0.19
0.18
CD
!!
t\I
0.17
of!
0
..
,g
0(
0.16
0.15
0.14
0.13
400 600
Wavelength (nm)
Figure 56.3 Electronic spectra of
MOTe complexes
BOO
146
NiYMDTC also exhibits transitions characteristic of octahedral complexes.
In the case of NiYMDTC the bands seen at at 15000cm-\ and 25000cm-\ could be
attributed to 3T\g(F) +- 3 A2g(F) and 3T\g(P) +- 3 A2g(F) transitions [18]. The charge
transfer transition is seen at 31060cm-\. The magnetic moment of 3.20 BM
observed for encapsulated Ni(U) complex also suggests an octahedral geometry.
Neat nickel dithiocarbamate complexes are reported to have square planar
geometry. The water molecules in the zeolite lattice or the oxide ions of the
framework are likely to involve in the octahedral geometry for Ni(n) ion.
CuYMDTC exhibits transitions at 14900cm-\ and 17000cm-\ in accordance
with B
2g
+-B\g and A\g +-B\g transitions [18, 23] indicating square planar
geometry around the Cu(U) ion. The high-energy band at 28570cm-\ can be
attributed to charge transfer transition. The magnetic moment of l.82 BM of
eu YMDTC agrees with the square planar structure.
SB.4.7. FTffi spectra
The important infrared spectral data of MDTC complexes are presented in
Table 5B.5 and the spectra are shown in Figure 5B.4. The absorption at 1461cm-\
in the spectrum of the ligand MDTC arises due to vN=CS
2
vibration. This vN=CS
2
is seen in the Co(III), Ni(U), and Cu(U) complexes in the region 1485cm-\ to
1500cm-\. The positive shift of this absorption may be an indication of the
interaction of the chelated sulphur atoms with the metal. [20]. The vC-S stretching
vibration seen in the spectrum of the ligand in the region 980-1 060cm-\ might have
shifted to lower frequency on complexation. But these bands are not observed in
the spectra of Co and Ni zeolite encapsulated complexes as these have been
masked by the broad zeolite absorption around 1000cm-\. This is seen at
950cm-\ in the spectrum of the Cu complex.
147
Table 5B.5 Infrared spectral data
............. ou. ........ "" ....................................................................... "', ..............................................................................................................................................................
MDTC CoYMDTC NiYMDTC CuYMDTC
-[ -[
cm-[
-[
cm cm cm
.._-_ .._ .._ ......._--------------
416 409
483 469 460 488
540 503
580 570 570 570
637 621
713 700 711
980 950
1023 1030 1020 1020
1112
1215 1234
1261 1267 1264
1322
1346 1354
1385 1385
1418
1461 1485 1500 1500
1625
1640 1635 1630
2112
2335 2336 2335 2336
2361 2361 2361 2361
2756
2852
2967 2970
3196
3279
3364 3475 3475 3472
.---
.... _ ...... __ ...
................_ .................."""""-,-
148
(iv)
3500 3000 z ~ 1000
Figure SB.4 Infrared spectra
i) MDTC and ii) CoYMDTC iii) NiYMDTC, iv) CuYMDTC
149
SB.4.8. EPR spectrum
Figure 5B.5 shows the EPR spectrum of CuYMDTC at liquid nitrogen
temperature. The calculated EPR parameters are given in Table 5B.6. The EPR
spectrum at room temperature is broad, and that at liquid nitrogen temperature
shows hyperfine splittings. The gll value of2.31 suggests a square planar structure.
Table SB.6 EPR parameters
2.31
109.2cm
2.02
0.9445
l.84 BM
P
0.l509
The ratio gllA
11
is 109.2 cm for CuYMTDC, suggesting a square planar
environment around Cu (II) ion. The value of a? is 0.9495, suggesting ionic
environment for the Cu
2
+ ions in the zeolite cavity. The magnetic moment
calculated using EPR parameters is 1.84 BM, which is in agreement with the value
obtained by Gouy method. All these suggest a square planar structure for the
zeolite encapsulated copper morpholinecarbodithioate complex.
150
_ ........................................ _ ...................................................... _ ...................................... u ___ ............ ................................. .
~ :
~
?
I
i
:.. ............................................................. - .................................... ...... ,. ....................... -.-.................................... ----....... ..
3100
Figure SB.S EPR spectrum of Cu YMDTC complex
SB.4.9. TG analysis
TG curves are shown in Figure 5B.6 and the TG data are tabulated in Table
5B.7.
Table SB.7 Thermal decomposition data
CoYMDTC NiYMDTC CuYMDTC
Temp range % Mass Temp range % Mass Temp range % Mass
(0C)
loss Cc)
loss
( QC)
loss
31 - 100 5.80 30 - 101 5.67 30 -100 5.61
100- 400 12.52 101 - 400 12.36 100- 460 12.51
400-799 0.36 400-798 0.34 460-798 0.44
- - - - - - - - ~
------
~ ......-----.
151
(iii)
(ii)
(i)
30
800
Figure SB.6 TG curves ofMDTC complexes
(i) CoYMDTC, (ii) NiYMDTC, and (iii) CuYMDTC
All the complexes lose free water in the temperature range 30 - 101C.
The mass loss in the temperature range 101 - 400 C corresponds to the loss of
coordinated water and the partial decomposition of the ligand. The mass loss
corresponding to this stage is - 12 %. The very small mass loss of - 0.34 % - 0.44
% in the temperature range 400 - 798C in each case can be due to the partial
decomposition of the complex as well as that of the zeolite.
152
References
1. D.F. Lewis, SJ. Lippard, and J.A Zubieta, Inorg. Chem. 11 (1972) 823.
2. G.N. Schrauzer, Transition Met. Chem. 4 (1968) 299.
3. L.H. Pignolet and S. H. Wheeler,Inorg. Chem. 19 (1980) 935.
4. J.R Hendrickson, J.M. Hope, and RL. Martin, J. Chem. Soc. Dalton. Trans.
(1976) 2032.
5. KW. Given, B.M. Mattson, M.F. McGuiggan, G.L. Miessler, and L.H.
Pignolet, J. Am. Chem. Soc. 99 (1977) 4855.
6. G.L. Miessler and L.H. Pignolet, Inorg. Chem. 18 (1979) 210.
7. N.M. Coleman, J.R Shelton and J.L. Koering, Ind. Engng. Chem. Prod. Res.
Develop. 13 (1974) 154.
8. D. Coucouvanis, Prog. Inorg. Chem. 11 (1970) 233.
9. D. Coucouvanis, Prog. Inorg. Chem. 26 (1979) 301.
10. R Eisenberg, Prog. Inorg. Chem. 12 (1970) 295.
11. J. Willemse, J.A Cras, JJ. Steggerda, and c.P. Keijzers, Stroct. Bonding
(Berlin) 28 (1976) 83.
12. K Tsuchiya, M. Yoshizumi, H. Houchi, and RP. Mason, J. Bioi. Chem. 275,
3 (2000) 1551.
13. A Godelitsas, D. Charistos, J. Dwyer, C. Tsipis, A Filippidis, A
Hatzidimitriou, and E. Pavlidou, Microporous and Mesoporolls Materials 33
(1999) 77.
14. N.N. Greenwood and A Eamshaw, Chemistry of Elements (1 st edn.),
Pergamon Press, Oxford, 1984. Rpt.1989.
15. F.A Cotton and G. Wilkinson, Advanced Inorganic Chemistry (5
th
edn .),
John Wiley and Sons, New York, 1988.
16. P. Gallezot, YB. Taarit, and B. Imolik, J. Catal. 26 (1972) 205.
17. H.D. Simpson and H. Steinfink, J. Am. Chem. Soc. 91 (1969) 6229.
18. AB.P. Lever, Inorganic Electronic Spectroscopy, Elsevier, Amsterdam,
1968.
19. KKM. Yusuff and EJ. Mathew, Synth. React. Inorg. Met. -Org. Chem. 19,1
(1989) 15.
20 L.M. Engelhardt, P.c. Healy, RM. Shephard, B.W. Skelton, and AH. White,
Inorg. Chem. 27 (1988) 2371.
21. KA Jensen and L. Henriksen,Acta. Chem. Scand. 22 (1968) 1107.
22. KS. Arulsamy, RF.N. Ashok and UC. Agarwala, Ind. J. Chem. 23A (1984)
127.
23. N.T. Madhu and P.K Radhakrishnan, Synth. React. Inorg. Met.-Org. Chem.
31,2 (2001) 315.
24. U Sakaguchi and AW. Addison, J. Chem. Soc., Dalton Trans. (1979) 600.
153
Chapter 6
Zeolite encapsulated palladium complexes
Abstract
Palladium compounds have been used as catalysts in a number of oxidation reactions
because of their activity even at very low concentrations. The possibility of using zeolite Y
encapsulated palladium complexes as oxidation catalysts was explored in the present study. Zeolite
encapsulated palladium compounds of DMG, MDTC and SALSC were synthesized using the
flexible ligand method. These compounds were characterized using chemical analysis, XRD, SEM,
surface area and pore volume analysis, DRS, FTIR, and TG measurements.
6.1. Introduction
Palladium has gained importance as an active component in catalytic reactions in
recent years. The most common use of palladium-based catalysts has been in
hydrogenation reactions. In recent times they are increasingly being used in C-C
coupling reactions of aromatic and vinylic systems [1-3]. Some compounds of
palladium in low oxidation states can also act as good oxidation catalysts. The
method for the oxidation of ethylene to acetaldehyde catalyzed by palladium(Il)
compounds, commonly referred to as the Wacker process, was reported by Smidt
and co workers [4-7]. The commercial success of the Wacker process provided an
enormous stimulus for further studies of palladium complexes as homogeneous
catalysts [8]. But Wacker process has some major disadvantages. The low pH of
the solution, high chloride concentration, and the formation of chlorinated
byproducts caused serious technical and environmental problems.
Homogeneous palladium catalysts are prone to a number of deactivation
reactions. One way out was to heterogenize the soluble catalyst. A number of
supports were used by Viswanathan for the study of palladium [9]. It has been
reported that the nature of the support used for palladium catalyst has a role to play
in modifying the selectivity towards useful products [10]. Supported palladium is
an efficient automotive catalyst and has been studied using a wide variety of
methods [11,12]. Supported palladium catalyst is found to be resistant to the action
of poison, especially with nitrogen donor poisons [13].
Palladium, like other precious metals and their compounds, is used in many
industrial processes in spite of its cost. These compounds are preferred because of
their high activity even at low concentration level, and their high specificity. The
possibility of using zeolite Y encapsulated palladium complexes as oxidation
catalysts was explored in the present study.
155
Complexes with nitrogen, oxygen, and sulphur donor ligands capable of
forming stable four, five, or six membered ring compounds with metal are well
known. Ligands can not only stabilize a metal complex catalyst but also
fundamentally change its selectivity and activity. Dithiocarbamates,
dimethylglyoximes, and semicarbazones can form complexes of palladium with
four, five, or six membered rings. Oithiocarbamates yield two 4-membered ring
compounds and dimethylglyoxime produces two 5-membered ring compounds
whereas salicylaldehyde semicarbazone produces one 6-membered and one 5-
membered rings.
In the present investigation, palladium complexes of DMG, SALSC and
MOTC were encapsulated in zeolite for spectral and catalytic studies. The
oxidative property shown by Pd(n) complexes was found to be retained on
encapsulation. Though the encapsulation of these complexes involves only
physical entrapment, selectivity, activity and recyclability were found to increase
several times. The results of our studies on these complexes are presented in this
chapter. The structures of SALSC and MOTC are presented in Figure 3B.1
(Chapter 3) and in Figure 5.1 (Chapter 5) respectively. The structure of
dimethylglyoxime is shown below.
Figure 6.1 Structure of dimethylglyoxime (DMG)
156
6.2. Experimental
Details regarding the synthesis and purification of the ligands
salicylaldehydesemicarbazone (SALSC), dimethylglyoxime (DMG), and
morpholinecarbodithioate (MDTC) are given in Chapter 2. Palladium exchanged
zeolite was prepared according to the procedure given in Chapter 2.
6.2.1. Synthesis of zeolite encapsulated palladium complexes
Palladium exchanged zeolite (3.0g) was treated with dimethylglyoxime so
that the ligand to metal ratio was greater than 2. The mixture was heated in a
sealed ampule at 120 QC for 18 hours. Impure PdYDMG was soxhlet extracted
using methanol. A similar method was followed for the preparation of PdYSALSC
(ligand to metal ratio slightly higher than 1) with the temperature at 200 QC for 6
hours. The impure compound was purified by soxhlet extraction using methanol
and chloroform.
PdYMDTC was prepared by adding PdY (3.0g) to a solution of MDTC in
water (ligand to metal ratio slightly higher than 2). It was refluxed for 5 hours.
After washing with water, surface species were removed using soxhlet extraction
with methanol. All the complexes were back exchanged with NaCI and washed to
remove chloride ions. They were dried at 110 QC for 2 hours and stored in vacuum
over anhydrous calcium chloride.
6.3. Characterization techniques
Details regarding the analytical methods and characterization techniques
are given in Chapter 2.
157
6.4. Results and Discussion
6.4.1. Chemical analysis
The results of chemical analysis are given in Table 6.1. The Si/AI ratio of
2.4 for the complexes is in the range expected for zeolite Y. The metal content is
low. All the metal ions present seem to be complexed. The good catalytic activity
can be the result of low metal ion concentration with high dispersion.
Table 6.1 Analytical data
Sample
PdYDMG
PdYSALSC
PdYMDTC
%Si
18.50
18.49
18.54
6.4.2. X-ray diffraction patterns
%Al
7.65
7.70
7.80
%Na
5.72
5.80
5.65
%C
0.09
0.08
0.03
%Pd
0.10
0.09
0.03
The XRD patterns of NaY, PdYDMG, PdYSALSC, and PdYMDTC are
given in Figure 6.2 .The XRD patterns of the palladium compounds show no peaks
corresponding to any palladium species. The main peaks are associated with the
crystallinity of the zeolite support. X-ray diffraction patterns indicate the good
crystallinity of these complexes. The XRD patterns of all these complexes were
similar to those of NaY, showing that the zeolite framework has remained
unaffected during the synthesis of the complexes.
158
5
10 15 20 25 30 40
Figure 6.2 XRD patterns of (i) NaY, (ii) PdYDMG, (iii) PdYSALSC
and (iv) PdYMDTC
159
6.4.3. Scanning electron microscopy
(i)
Figure 6.3 SEM
i) Before soxhlet extraction
ii) After soxhlet extraction
160
The scanning electron micrographs of the palladium dimethylgloxime
complex taken before and after soxhlet extraction indicate the removal of surface
adsorbed species. This is clear from the figure.
6.4.4. Surface area and pore volume
The BET surface area and pore volume data are given in Table 6.2. The
encapsulation of PdYDMG, PdYSALSC, and PdYMDTC complexes inside the
zeolite cavity is indicated by the lowering of surface area and pore volume
compared to the metal exchanged zeolite.
Table 6.2 Surface area and pore volume data
Surface area (m2/g) Pore volume (mllg)
Sample
PdY PdYL % loss PdY PdYL % loss
PdYDMG 500 220 56.0 0.2898 0.1300 55.0
PdYMDTC 500 180 64.0 0.2898 0.1021 64.7
PdYSALSC 500 228 54.4 0.2898 0.1352 53.7
6.4.5. Magnetic moment
Magnetic moment values calculated from room temperature magnetic
susceptibility methods suggest that all the zeolite Y encapsulated palladium
complexes are diamagnetic. This suggests square planar geometry for the
palladium complexes.
161
6.4.6. Diffuse reflectance spectra
The diffuse reflectance spectral data are given in Table 6.3 and the spectra
are presented in Figure 6.4. Similar transitions are seen in all the zeolite Y
encapsulated complexes arising form the' A'g level. The band at 21370cm
o
'seen in
all the complexes is due to 'A2g -'A,g transition. The band seen at 26600cm
o
' in
the case of PdYDMG and that at 27000cm
o
' in the case of PdYMDTC and
PdYSALSC are due to 'Eg - 'A,s transition. The band at 33000cm
o
' in all the
complexes is the result of 'B'g - 'A,g. These are characteristic of square planar
geometry, and indicate square planar geometry for palladium complexes [14].
Room temperature magnetic measurement further supports this inference.
Table 63 Electronic spectral data
Sample Assignment
PdYDMG
21370 'A '
2g- A'g
26600
'Eg- 'A,g
33000 'B'g-'A,g
21370
' ,
A2g- A'g
27000 'Eg-'A,g
33000 'B'g-'A,g
PdYSALSC
21370
'A2g- 'A,g
27000 'Eg-'A,g
33000 'B'g-'A,g
PdYIvIDTC
162
0 . 5 r ~
0 . 0
0.35
'"
!:!
.e 0.30
o
III
s:J
<
0.25
0.20
I PdYMDTC I
o. , 5 '-"'---..... 00-----''----6 .... 00----''---'80-:''0 ....
0.8
0.7
~ 06

o
III
s:J
<
0.5
o.
Wavelength (nm)
I PdYDMG I
O. 3 '-"'---' .... 00-----''----6 .... 00--....... ---'80'-0 ....
Wavelength (nm)
05
PdYSALSC
'00 600 800
Wavelength (nm)
Figure 6.4 Electronic spectra of
palladium complexes
163
6.4.7. FTIR spectra
A comparative study of the spectra of the Iigands and the complexes was
made. The IR spectra of DMG and PdYDMG are shown in Figure 6.5, those of
SALSC and PdYSALSC in Figure 6.6, and of MOTC and PdYMOTC in Figure
6.7. The spectral data are shown in Table 6.4. The IR spectra of DMG and its
complexes have been studied extensively by Shpiro et al. [15]. In the spectrum of
ligand DMG, two absorptions each have been reported for C=N and N-O
vibrations. These two bands can be due to unequal C=N and N-O linkages. In the
case of free DMG ligand, the bands which occur at 1440cm-
1
and 1594cm-
1
are due
to C=N stretching vibration and those at 980cm-
1
and 1143cm-
1
are due to N-O
stretching frequencies_ In the spectrum of the complex, the band at 1530cm-
1
can
be due to the C=N stretching vibration. The shifting of the band at 1594cm-
1
to
lower frequency suggests coordination through nitrogen. The peak due to N-O
stretching vibration is not seen in the spectrum of the complex as it is masked by
the broad band of zeolite.
In the spectrum of the ligand salicylaldehyde semicarbazone, the
absorption at 3460cm-
1
arises as a result of vO-H and vNH
2
vibrations [16]. The
ligand band at 1620cm-
1
due to vc=o in the ligand is shifted to 1600 cm-
I
in the
complex. The vC=N stretch at 1485 cm-
I
is shifted to lower frequency indicating
that coordination has occurred also through nitrogen. The vc-o of the phenolic
group is masked by absorption of the zeolite. The absorption bands of vNH and
8NH
2
, which are expected to remain unaltered in the spectrum of the complexes, are
masked by the absorption of water.
In the dithiocarbamate, the bonding to the central metal is through the
sulphur atoms of the ligand. The VN=CS
z
absorption of morpholinecarbodithioate
is seen at 1461 cm -I. In the spectrum of the complex, this band is observed as a
very weak band at 1493cm-
l
. This positive shift indicates the interaction of
chelated sulphur atoms with the metal ion [17].
164
Table 6.4 Infrared spectral data
OMG PdYOMG SALSC PdYSALSC MOTC PdYMOTC
cm-! cm-! cm-! cm-! cm-! cm-!
466 458 482 476
561 540 580
705 618 678 617 637 618
753 722 750 773 868
892
902 894 980
980 980 1023 1004
1038 1030 1032 1036
1063
1143 1110 1112 1097
1365 1150 1215
1440 1201 1261
1594 1530 1273 1300
1637 1353 1346
1941 1384 1380 1385
2369 2369 1485 1415 1418
3214 1620 1600 1461 1492
1640
2367 2368 1625 1640
3687 3465 2829 2832 2112
3416 3460 2335 2335
2361 2361
2756
2852
3196
3279
3364 3480
165
4000 35(10 iOOO 2500 2000 1500
Wavenumber cm-!
1000
Figure 6.5 IR of ligand DMG and DMG complex
(i)
500
166
(i)
4000 3000 2500 2DOO 1500 1000 SOD
Wavenurnber cm-
1
Figure 6.6 IR of ligand SALSC and SALSC complex
167
(ii)
Figure 6.7 IR spectra
i) MDTC and ii) PdYMDTC complex
168
6.4.8. TG analysis
The TG curves of PdYDMG were recorded in an atmosphere of nitrogen
from room temperature to 800C at a heating rate of lOO/min. Figure 6.7 shows
the TG curve of PdYDMG. Thenno analytical data for the complexes are
presented in Table 6.5.
3.2
3.1
3.0
Cl 2.9
..
1:
01
2.8
'Q;

2.7
2.6
2.5
Table 6.5 Thennal decomposition data
0
PdYDMG
Temp Range % Mass
(0C)
loss
40 -120 8.0
120 - 400 6.3
400 -780 4.4
\
\
I -PdYDMG I
\
\
100 200 300 400 500 600 700 800
Temperature ('C)
Figure 6.8 TG curve of PdYDMG
169
The percentage mass loss of PdYDMG at different temperature ranges is
given in the table. The loss of free and coordinated water occurs up to 125 QC. The
decomposition of the ligand and the partial decomposition of zeolite occur as a
continuous process. We get only a qualitative idea from this curve.
170
References
1. 1. Tsuji, Palladium Reagents and Catalysts, Wiley, New York, 1996.
2. I.P. Beletskaya and A.V. Cheprakov, Chem. Rev. 100 (2000) 3009.
3. Jairton Dupont, Roberto F. de Souza, and Paulo A.Z. Suarez, Chem. Rev.
102 (2002) 3667.
4. 1. Smidt, Chem. Int. (London) (1962) 54.
5. H. Arai, T. Yamashiro, T. Kubo, and H. Tominago Bull. Jpn. Pet. Inst. 18
(1976) 39.
6. T .Kubota, F. Kumada, H. Tominago, and T. Kanugi, Int. Chem. Eng. 13
(1973) 539.
7. 1. Smidt, W. Hafner, R. Jira, J. Sedlmeier, and A. Sabel, Angew. Chem. Int.
Ed. 1 (1962) 80.
8. P. M. MaitIis, The Organic Chemistry of Palladium. Vols. I and 2,
Academic Press, New York, 1971.
9. N. Mahata, K.V. Raghavan, and V. Viswanathan, Bulletin of the Catalysis
Society of India 9,1-2 (1999) 75.
10. S. Chandra Shekar, A. Venugopal, K.S. Rama Rao, P.S. Sai Prasad, R.
Srinivas, and P. Kanta Rao, in Recent Advances in Basic and Applied
Aspects of Industrial Catalysis (Studies in Surface Science and Catalysis
113), T.S.R. Prasad Rao and G. Murali Dhar (Eds.), Elsevier Science, 1998,
391.
11. A.V. Kucherov and M. Shelef, Catal. Lell. 75, 1-2 (2001) 19.
12. A. Fritz and V. Pitchon, Appl. Catal. B 13 (1997) 1.
13 C.e. Costa Augusto, J. Luiz, Zotin, and Arnaldo da Costa Faro Jr., Catal.
Let!. 75, 1-2 (2001) 37.
14. A.B.P. Lever, Inorganic Electronic Spectroscopy, Elsevier, Amsterdam,
1968.
15. E.S. Shpiro, G.V. Antoshin, O.P. Tkachenko, S.V. Gudkov, B.V.
Romanovsky, and Kh.M. Minachev, Stud. Surf Sci. Catal. 18 (1984) 31.
16. K. Nakamoto, Infrared Spectra of Inorganic and Coordination Compounds.
John Wiley, New York, 1970.
17. L.M. Engelhardt, P.e. Healy, R.M. Shephard, B.W. Skelton, and A.H. White,
Inorg. Chem. 27 (1988) 2371.
171
Chapter 7
Catalytic activity of zeolite encapsulated complexes in the
decomposition of hydrogen peroxide
Abstract
Hydrogen peroxide participates in a variety of chemical reactions. The activity of zeolite Y
supported transition metal complexes in the decomposition of hydrogen peroxide has been studied
in detail with a view to screening the catalytic activity of the complexes in oxidation reactions. The
reaction was studied by measuring the volume of oxygen evolved using a gas burette. The neat
complex and ion exchanged zeolite were also screened for the catalytic activity. Among the various
catalysts investigated, PdYSALSC was found to be the most active catalyst, with Cu YMDTC
coming next. The catalytic activity was studied by varying parameters like the volume of H
2
0
2
,
amount of catalyst, and solvent polarity. The effects of the addition of pyridine and the use of
recycled catalysts were also studied.
7.1. Introduction
Zeolite encapsulated complexes are known to exhibit similar/ better catalytic
behaviour compared to homogeneous complex catalysts. The study of catalysis by
encapsulated complexes started with the heterogenization of the complexes of
known catalytic activity in solution. In the encapsulated state it is possible to
obtain catalytically active centres that are unusual for complexes in solution. The
chief advantage of encapsulated complexes is the ease in separation from the
reaction mixture, which reduces cost of recovery. Hydrogen peroxide is
industrially and chemically important because of its use in a variety of processes
[1-3]. The decomposition of hydrogen peroxide represented by the reaction
2 H
2
0
2
-+ 2 H
2
0 + O
2
is considered a standard reaction to determine the catalytic activity of metal
complexes [4]. The above reaction can be catalyzed by several transition metal
ions and transition metal complexes, provided the coordination sphere is accessible
to the H
2
0
2
or HOO- ion [5-6]. The efficiency of such systems depends to a large
extent on the structural and electronic properties of encapsulated complexes. The
zeolite support plays an important role in the dispersion of metal complex and in
increasing the stability of the catalyst in addition to influencing catalytic
properties. Considering the significance of oxidation reactions using hydrogen
peroxide as oxidant for the synthesis of fine chemicals, pharmaceuticals, speciality
products and so on, it was thought important to study the decomposition of
hydrogen peroxide using the different encapsulated complexes.
7.2. Experimental
7.2.1. Materials
Zeolite encapsulated complexes ofSPO, SPP, SALSC, SB, DTC and DMG
were used for catalytic studies. The encapsulated complexes were prepared
according to the procedures reported in Chapters 3, 4, 5, and 6. Neat complexes of
173
most catalytically active encapsulated complexes were also prepared according to
reported procedures for comparison of the catalytic activity. The various reagents
used are mentioned in Chapter 2.
7.2.2. Procedure for the decomposition of hydrogen peroxide
All the catalysts were activated by heating to 120 QC for two hours. A
plastic float containing the sample catalyst was placed over hydrogen peroxide
taken in a reaction flask of 50ml capacity containing a magnetic paddle for
stirring. The quantities of different reagents used are specified in each table. A gas
burette was used for collecting the oxygen evolved. The gas burette was filled with
20 % NaCI solution acidified with dilute hydrochloric acid and coloured using
methyl red. The solution levels in both arms of the gas burette were equalised at
zero before sealing the reaction flask. The magnetic stirrer was switched on,
starting a stopwatch simultaneously. The catalyst was dispersed in hydrogen
peroxide solution by magnetic stirring. As the reaction proceeded, oxygen was
liberated and it got collected in the right arm of the gas burette. The volume of
oxygen liberated was noted after equalizing the solution levels in both arms at
intervals of 5 minutes. The experiment was continued for 30 minutes/until 1 ami of
gas was collected in the burette. The reaction was carried out under different
conditions. The conditions specific to each test are given along the results. The
catalyst was washed with acetone and heated at 120 QC for 2 hours and reused.
7.3. Results
7.3.1. Screening studies
A comparison of the catalytic activities of SALSC, DTC, SB, sPa, SPP
and DMG encapsulated complexes in the decomposition of 30 %(w/v) H
2
0
2
at
174
Table 7.1 A comparison of the catalytic activity of various solid catalysts in the
decomposition of 30% (w/v) H
2
0
2
at room temperature
SALSC
FeY
CoY
0.2
NiY
0.15
CuY
0.9
PdY
9
9
B
7
6
Volurneof
5
oxygen (rnl) 4
3
2
0-
Volume of oxygen evolved (ml)
MDTC SB SPO
0.1 0.1
0.3 0.6
0.2 0.15
4.1 1.2 0.6
0.9
Hydrogen peroxide decomposition
DSALSC.
!:IDTC
I:iSB
8SPO
IlISPP
.DMG
FeY CoY NiY
Metal Ions
CuY
SPP
0.1
0.2
0.1
0.3
PdY
DMG
0.1
Figure 7.1 Volume of oxygen liberated using different catalysts
Reaction condition
Weight of catalyst : 10 mg
Temperature : 30C
Duration : 30 min
Vol. 0[30% H
2
0
2
: 10 ml
Vol. of H
2
0 : 10 ml
175
Table 7.2 A comparison of the catalytic activities ofPdYSALSC,
CuYMDTC, and CuYSB for H
2
0
2
decomposition
Time (min) Volume of O
2
liberated (ml)
o
5
10
15
20
25
30
g
Q)
E
"
~
10
8
6
4
2
0
PdYSALSC
0.1
2.5
4.45
6.55
9.0
-e- PdYSALSC
--CuYMDTC
--CuYSB
CuYMDTC
0.1
0.5
1.1
1.7
2.5
3.2
4.1
e
/
0/0
/ ~ ~
e ~
e ~ .-.
.-?-. .-... .-
e
o 5 10 15
Time (min)
20 25 30
CuYSB
0.1
0.15
0.25
0.5
0.65
0.95
1.2
Figure 7.2 Volume of oxygen evolved versus time
Reaction condition
Weight of catalyst : 10 mg
Temperature : 30 QC
Duration : 30 min
Vol. of30% H
2
0
2
: 10 ml
Vol. of H
2
0 : 10 ml
176
room temperature is given in Table 7.1. A bar chart diagram drawn with the
volume of oxygen liberated against the different complexes is shown in Figure 7.1.
The reaction was continued for 30 minutes or until 10 ml of oxygen was evolved.
Figure 7.2 shows the volume of oxygen against time for the three complexes
PdYSALSC, CuYMDTC and CuYSB; and Table 7.2 gives a comparison of the
catalytic activities ofPdYSALSC, CuYMDTC and CuYSB for hydrogen peroxide
decomposition. The decomposition data show that PdYSALSC is the most active
catalyst. CuYMDTC is found to be moderately active, while CuYSB is weakly
active. There is only minor difference in the activities of CuYSALSC, PdYMDTC
and CuYSPO. The catalytic activity ofPdYMDTC is significantly less than that of
PdYSALSC even though the metal ions are the same. The activity of Cu YMDTC
is much higher compared to that of CuYSB. This shows that catalysis greatly
depends on the type of chelating ligand employed [7].
7.3.2. Blank run
The decomposition reaction was studied under conditions identical with
those of screening experiments, but without adding the catalysts. There was no
measurable decomposition of hydrogen peroxide for up to 30 minutes. The
decomposition reaction was further studied using various metal exchanged zeolites
under conditions identical to those of the above-mentioned blank. The volume of
oxygen liberated with each metal exchanged zeolite is shown in Table 7.3. Figure
7.3 is a plot of volume of oxygen liberated for the different metal exchanged
zeolites. Of all the metal exchanged zeolites, PdY showed maximum activity. But
this was low compared with the volume of oxygen liberated using PdYSALSC.
The catalytic activity of PdYSALSC was approximately five times greater than the
activity of PdY. The volume of oxygen liberated with other metal exchanged
zeolites was negligibly low.
177
Table. 7.3 Comparison of the catalytic activity of various ion exchanged zeolites
in the decomposition of 30 % H
2
0
2
at room temperature
Metal exchanged zeolite Volume of oxygen
(ml)
2.5-
c:
2
>-
___ 1.5-
'O"E
Q)'-'" 1
E
:l
"0 0.5
>
o
FeY
CoY
NiY
CuY
PdY
0.1
0.2
0.1
0.3
2.5
FeY CoY NiY CuY PdY
Ion exchanged zeolite
Figure 7.3 Ion exchanged zeolite versus volume of oxygen
Reaction condition
Weight of catalyst : 10 mg
Temperature : 30C
Duration : 30 min
Vol. of30% H
2
0
2
: 10 ml
Vol. of H
2
0 : 10 ml
178
The decomposition of hydrogen peroxide was also studied under identical
conditions using neat Cu(MDTC)2, Pd(SALSC)Cl and CU(SB)2Ch . Neat
CU(SB)2Ch showed no activity. Oxygen evolution occurred with neat Cu(MDTC)2
and neat Pd(SALSC)CI. But the activity observed with neat Cu(MDTC)2 and neat
Pd(SALSC)Cl was very small compared to that shown by encapsulated complexes.
7.3.3. Factors influencing the decomposition of hydrogen peroxide
In order to understand the catalytic nature of PdYSALSC, CuYMDTC and
Cu YSB on the decomposition of hydrogen peroxide, the various factors which
influence the decomposition ofH
2
02 were studied. The factors studied were:
volume ofH20 2
amount of catalyst
solvent polarity of the reacting solute
addition of pyridine
addition of alkali
using recycled catalysts
7.3.3.1. Influence of the volume of hydrogen peroxide
The effect of H
2
0
2
1H
2
0 volume ratio on the decomposition of H202 with
PdYSALSC, CuYMDTC, and CuYSB was studied. The results are given in Table
7.4. The total volume of the mixture of H
2
0
2
and water was kept constant. The
reaction conditions were identical to those in the earlier cases. The effect was
studied by varying the amount of hydrogen peroxide. It was found that the volume
of oxygen evolved also increased with the increase in the volume of H
2
0
2
. The
same effect was observed in all the three cases.
179
Table 7.4 Effect ofH
2
0
2
1H
2
0 volume ratios on the decomposition ofH
2
0
2
with PdYSALSC, CuYMDTC, and CuYSB as catalysts
H20
2
1H
2
O Volume of oxygen Cml)
Cml) PdYSALSC CuYMDTC CuYSB
5: 15 7.2
10: 10 9
3.2 0.95
15:5 9
4.15 2.1
20:0
6 2.75
Reaction condition
Weight of catalyst: 10 mg
Temperature : 30 QC
Duration : 25 min
7.3.3.2. Influence of the amount of catalyst
The effect of varying the amount of catalyst was studied in the case of
PdYSALSC, CuYMDTC, and CuYSB. The mass of the catalyst was varied from 5
mg to 15 mg. The proportion of the mixture of H
2
0
2
and H
2
0 is kept constant at
10 ml: 10 ml. The reaction was carried out under identical conditions as that ofthe
screening test. The oxygen evolution increased with increase in the amount of the
catalyst in the case of PdYSALSC and Cu YMDTC. The results are given in Table
7.5, and Figure7.4 shows the volume of oxygen evolved with amount of catalyst.
180
Table 7.5 Effect of variation in the amount of catalyst on the decomposition of
H
2
0
2
Weight of Catalyst (mg)
.
III
E
"
"0
>
5
10
15
11
10
9
8
7
6
5
4
3
2
0
e
,,-
lI!
4 6
Volume of oxygen (ml)
PdYSALSC CuYMDTC
3.2
6.7
10.0
-e- PdYSALSC
-,,- CuYMDTC
-lIE-CuYSB
e
1.8
2.5
4.0
CuYSB
e
0.4
0.7
0.7
- ~
lIE lI!
B 10 12 14 16
Weight (mg)
Figure 7.4 Amount of catalyst versus volume of oxygen
Reaction condition
Temperature : 30C
Duration : 20 min
Volume of30% H202: lOml
Volume of H
2
0 : 10 ml
181
7.3.3.3. Influence of solvent polarity of the reaction solution
The effect of the solvent polarity of the reaction solution was studied. The
reaction was studied by adding methanol to the reaction mixture (by changing the
volume of water and methanol in such a manner as to keep the total volume
constant). The dielectric constant of the medium was calculated assuming a linear
relationship between the individual dielectric constants and the composition of the
solvent. Thus if C is the resultant dielectric constant, Cl and C2 the dielectric
constants of water and methanol, and VI and V
2
their volumes employed
respectively, C may be given as
V
where V is the total volume of the reaction mixture.
The values of Cl and C2 are 78.54 and 32.63 respectively. The results obtained in
the case of PdYSALSC is given in Table7.6, with CuYMDTC in Table 7.7 and
CuYSB in Table 7.8. Decomposition of hydrogen peroxide was found to be
slowed by the addition of methanol in all the three cases.
Table 7.6 Effect of solvent polarity on the reaction solution with PdYSALSC
H
2
0: CH
3
0H (ml) Volume of oxygen (ml)
Reaction condition
Weight of Catalyst: 10 mg
Temperature : 30C
Volume ofH
2
0
2
: 10 ml
Time : 25 min
10:0 9
9: 1 6.9
8: 2 5.5
182
Table 7.7 Effect of the solvent polarity on the reaction solution with CuYMDTC
Reaction condition
Weight of catalyst: 10 mg
Temperature : 30C
Volume ofH
2
0
2
: 10 ml
Time : 30 min
10:0
9: 1
8:2
Volume of oxygen
(ml)
4.1
0.5
0.3
Table 7.8 Effect of the solvent polarity on the reaction solution with CuYSB
Reaction condition
Weight of catalyst: 10 mg
Temperature : 30C
Volume ofH
2
0
2
: 10 ml
Time : 30 min
10:0
9: 1
8:2
Volume of oxygen
(ml)
1.2
1.2
0.9
183
7.3.4. Addition of pyridine
The effect of adding traces (two drops) of pyridine to the reaction mixture was
studied in the case ofPdYSALSC and CuYMDTC. The experiment was conducted
in two ways.
i) By adding pyridine to the catalyst before adding hydrogen peroxide to
it. No decomposition of hydrogen peroxide was observed.
ii) By adding the catalyst to the mixture of hydrogen peroxide and water and
then adding pyridine. Instead of deactivating the effect there was an
increase in the decomposition rate.
The results are presented in Tables 7.9 and 7.10.
Table 7.9 Effect of adding traces ofpyridine on using PdYSALSC catalyst
Time Volume of oxygen
(min) (ml)
o 0.1
5.0 3.0
10 9.7
Reaction condition
Temperature : 30 QC
Volume ofH202 : 10 ml
Volume of H
2
0 : 10 ml
Amount of pyridine: 2 drops
184
Table 7.10 Effect of adding traces ofpyridine on using CuYMDTC catalyst
Time
(min)
0
5
10
15
20
25
Reaction condition
Temperature : 30C
Volume ofH
2
0
2
: 10 ml
Volume of H
2
0 : 10 ml
Amount of pyridine: 2 drops
7.3.5. Recycled catalysts
Volume of oxygen
(ml)
0
1.7
3.45
5.05
6.7
7.3
The catalysts once used were washed with acetone, and dried in oven at
120C for two hours. The recycled catalyst was characterized using XRD, IR, and
electronic spectroscopy.
The XRD patterns of the recycled catalysts were similar to those of fresh
catalysts. This showed that the crystalline structure was retained even after the
regeneration process. The IR of the recycled sample resembled that of the fresh
sample. From this it was inferred that the complex remains unaffected in the cages
of zeolite Y. The electronic spectra of the used sample were similar to those of the
fresh sample, showing that there is no change in the structure even after recycling.
185
Table 7.11 Effect of using recycled catalyst on the decomposition on H
2
0
2
Time (min)
0
5
10
15
20
25
30
Reaction condition
Temperature : 30C
Volume ofH
2
0
2
: 10 ml
CuYSB
0.1
0.1
0.3
0.5
0.7
0.9
1.2
Volume of oxygen evolved (ml)
CuYMDTC PdYSALSC
0.1
0.3
0.7
1.2
1.8
2.4
3.1
Volume of H
2
0 :10 ml
Weight of catalyst: 10 mg
0.1
0.3
1.0
1.7
2.6
3.5
4.8
1 0 r ~
9
8
7
6
~ 5
'"
E 4
:>
g 3
2
o
o
-CuYSB
-A- CuYM DTC-recycled
-0- PdYSAlSCrecycled
_-PdYSAlSC
-Y-CuYMDTC
-x- CuYSB-recyled
10 15
time (m in)
20 25 30
Figure 7.5 Comparison of activity on using recycled catalyst and unused catalyst
in the decomposition of H
2
0
2
186
----------
80
------
-_. __ . -
C Recycled catalysts
, Fresh catalysts
I
Figure 7.6 Comparison of the retained activity and loss in activity of the
three catalysts
The electronic spectra of the used sample were similar to that of the fresh sample,
showing that there is no change in the structure even after recycling. The reaction
mixture after the first cycle was filtered to remove the catalyst. The resulting
solution was analyzed for metal content using atomic absorption spectroscopy. The
absence of metal ions in the reaction medium indicated that there was no leaching
out of the metal ions from the pores of the zeolite. This again proved that the
complexes were intact within the cages. The reaction was carried out under
identical conditions using recycled PdYSALSC, CuYMDTC, and CuYSB. The
results obtained are tabulated in Table 7.11 and presented graphically in Figure 7.5
and Figure 7.6. In the case of CuYSB the activity of the recycled catalyst was the
same as that of the unused catalyst, as shown in Figure 7.6. Recycled CuYMDTC
showed - 75% activity as compared to the fresh sample. The most deactivated
was the highly catalytic PdYSALSC.
187
7.4. Discussion
The decomposition of hydrogen peroxide was studied both using catalysts
and without using catalysts. Of the catalyzed reactions, those with encapsulated
complexes showed much higher activity compared to those with neat complexes
and ion exchanged zeolites. PdYSALSC was found to have a five times greater
activity than PdY at the same pH and other conditions. The activity of other ion
exchanged zeolites was negligible compared to that of encapsulated complexes.
The activity shown by neat complexes was also negligible compared to that of
encapsulated complexes. Though the encapsulated complex shows the general
characteristics of homogeneous catalysts, it also has the merits of heterogeneous
systems.
Many possible mechanisms have been proposed for the decomposition of
hydrogen peroxide. All these mechanisms involve the oxidation or reduction of the
substrate, i.e., the decomposition is facilitated by a redox cycle between the
oxidation states of the metal. One mechanism which can be applied in the case of
zeolite encapsulated complexes can be as follows:
2H
2
0
2
-+ 2HOO- + 2H+
2(LM)"+ + 2HOO- -+ 2(LM)(n-l}+ + 2HOO'
2HOO -+ O
2
+ H
2
0
2
2(LM)(n-l}+ + H
2
02 -+ 2(LM)n+ + 20H-
Eq. VII. 1
Eq. VII.2
Eq. VII.3
Eq. VII.4
Combining Eq.VII.2 and Eq. VII.4 and cancelling the catalyst term will give
2HOO- + H
2
0
2
-+ 2HOO' + 20H- Eq. VII.S
The redox potential of the active metal plays a vital role in the decomposition of
hydrogen peroxide because the catalytic activity of the metal ion depends on its
redox potential. The complex will have a modified potential, which will increase
the catalytic activity.
In another mechanism proposed, the decomposition reaction occurs in the
coordination sphere of the metal ion. This involves the formation of an
188
intermediate peroxo species. The following pathway may be suggested for this
mechanism:
(LM)n+ + H
2
0
2
- (LMH
2
0
2
)"+ Eq. VII.6
(LMH
2
0
2
)"+ - LM(HOO) (nI)+ + W Eq. VII.7
LM(HOO) (n-I)+ + H
2
0
2
_ LM(HOO)(H
2
0
2
) (n-I)+ Eq. VII.8
LM(HOO)(H202) (n-I)+ _ (LM)n+ + 02 + H
2
0 + OH- Eq. VII.9
The peroxo intermediate can also be formed by the action of HOO- generated
from the free H
2
0
2
molecules in solution.
_ LM(HOO) (n-I)+
Eq. VII. IQ
Eq. VII.ll
According to Sharma and Schubert, the peroxide entity in the transition state in Eq.
VII.7 can become highly reactive [8]. The accessibility of the coordination sphere
for H
2
0
2
or HOO- and the availability of at least two free vacant coordination
sites at the metal ion are the main requirements for the decomposition to proceed
via this mechanism.
Helmut Sigel has found that the decomposition rate is high with copper
polyammine chelates with vacant coordination sites whereas it is low when the
copper complexes are fully coordinated [9]. According to Sharma et aI., complexes
having less than two free sites on the metal ion are inactive [8]. This observation
can be related to the bidentate nature of peroxide anion or to the fact that the
remaining free positions on the intermediate will facilitate its reactions with the
second peroxide anion.
Decomposition data show that the encapsulated Pd(u) and Cu(u)
complexes are very active in the decomposition of hydrogen peroxide. These two
complexes have square planar structure. This provides vacant coordination sites
and facilitates the formation of peroxo intermediates which are highly reactive
species leading to the rapid decomposition of hydrogen peroxide. The mechanism
involving peroxo intermediate seems to be applicable in the case of encapsulated
Pd and Cu complexes, as these complexes showed a colour change on coming in
189
contact with hydrogen peroxide. This colour change is indicative of the formation
of intermediate peroxo species. Similar colour change has been found in the case
of CU(II) complexes by earlier workers [10].
A steady increase in pH occurs during the decomposition of hydrogen
peroxide, confirming the liberation of OH- ions as suggested by the above
mechanism. Such an increase in pH has been reported by earlier experimenters
[ 11].
The catalytic activities were found to be modified to a considerable extent
by ligands [12]. The effect of the ligands on the catalytic activity ofthe complexes
has been pointed out by Sharma and Schubert [8]. They have pointed out that the
reduction of Cu(n) coordinated to two nitrogen atoms is much faster than that of
aquo cupric ion. Wang suggests that the enhancement in catalytic activity is due to
the strain around the metal ion catalyst provoked by the ligand molecules [13].
Sharma and Schubert have pointed out that there is a trend towards higher catalytic
activity on increasing the ring size of the chelate [8]. This is in agreement with the
observation in the present study that SALSC complexes with six membered and
five membered rings exhibit maximum activity, and MDTC with two 4-membered
rings chelates with medium activity and SB without chelation has the lowest
activity. The higher activity of the 6-membered ring chelate can be related to the
greater ease with which it can remain coordinated to the metal ion during its
reversible oxidation - reduction [8].
The catalytic activity of the encapsulated Pd(n) and Cu(n) complexes were
much higher compared to that of the corresponding neat complexes. According to
one earlier report there is no significant difference in the kinetics of reactions using
supported catalyst and non-supported catalyst except in the enhancement of rate
[14]. This very much agrees with what we have found in our study. The
encapsulated systems were found to be much more active compared to neat
complexes. Enhanced activity with supported catalyst has been reported by Wilson
190
[15]. The enhanced activity of the supported system can be due to the formation of
a modified catalyst precursor involved in the rate-determining step [14], or due to
the greater stability of the encapsulated complex compared to the neat complex.
The decomposition reaction was studied by varying the volume of
hydrogen peroxide. The volume of oxygen evolved increased with the volume of
hydrogen peroxide. The rate of decomposition increased with the amount of the
catalyst in the case of PdYSALSC and CuYMDTC. Such an observation was
made by earlier workers [7,16]. This effect was negligible in the case ofCuYSB.
The addition of methanol was found to slow down the decomposition
reaction. This was observed with all the three catalysts. Since methanol is of lower
polarity than water, the decrease in reaction rate observed here must be because of
the decease in polarity.
The rate of a reaction is given by the equation
Ink = Inko - Z ... Zse
2
EdAskT
where:
k - rate constant
ko - rate constant in a solvent of infinite dielectric constant
ZAe, Zse - charges of the reactive species in the transition state respectively
E - dielectric constant of the solvent
dAB -diameter of the activated complex
k - Boltzmann's constant
T - absolute temperature
From the equation it is clear that the reactive species participating in the
formation of the transition state are either both positively charged or both
negatively charged for Ink to decrease. Pure complex is positively charged, and
H
2
0
2
and O
2
are neutral species. As there is a possibility of H
2
0
2
getting ionised
191
to H+ + HOO- , it might be the HOO- species that would be reacting with the
complex catalyst. The solvent effect studies indicate that the complex also might
be acting as a negative species, which can be formed inside the zeolite cavity due
to interaction with oxide ions of the lattice. Since the spectral indication is square
planar, the matrix can be distantly coordinated to the metal ion.
The addition of pyridine to the catalyst gave different results. The
experiment with the direct addition of pyridine to the catalyst before adding
hydrogen peroxide produced no decomposition of hydrogen peroxide. This can be
because the pyridine complex formed is less reactive than the original complex.
The experiment with the addition of catalysts to the mixture of hydrogen peroxide
and water followed by the addition of pyridine showed very high rate of
decomposition. This can be because pyridine acts as a base and decomposes
hydrogen peroxide as per the equation
H
2
0
2
- W + HOO-
HOO- ion is a better reactant than H202. The acceleration observed in the
present reaction can be due to the increased concentration of the H0
2
- ion. It has
also been reported that alkalinity has a strong influence on the activity of heavy
metal catalyzing species in the decomposition of hydrogen peroxide [17].
Recycled PdYSALSC and Cu YMDTC showed less activity compared to
fresh catalyst. Cu YSB showed no deactivation at all.
192
References
1. S. Tamagaki, M. Sasaki, and W. Tagaki, Bull. Chem. Soc. Jpn. 62 (1989)
153.
2. J.H. Buxendak, Adv. Catal. 4 (1952) 39.
3. H.M. Cota, T. Katan, M. Chin, and FJ. Schoenweis, Nature, 203 (1964)
1281.
4. R. Sreekala and K.K. Mohammed Yusuff, Ind J Chem., 34A (1995) 994.
5. H. Sigel,Angew. Chem. 81 (1969) 161.
6. H. Sigel, Angew. Chem., Int. Ed. Engl. 8 (1969) 167.
7. Isao Mochida, Akio Kato, and Tetsuro Suiyama, Bull. Chem. Soc. Jpn. 45
(1972) 2230.
8. V.S. Sharma and J. Schubert, Inorg. Chem. 10, 2 (1971) 251.
9. Helmut Sigel, Kurt Wyss, Beda E. Fischer, and Bernhard Prijs, Inorg. Chem.
18, 5 (1979) 1354.
10. R.V. Prasad and N.V. Thakkar, Ind J Chem. 33A (1994) 861.
11. Beena Pandit and Uma Chadusama, in Recent Advances in Basic and Applied
Aspects of Industrial Catalysis (Studies in Surface Science and Catalysis
113), T.S.R. Prasada Rao and G. Murali Dhar (Eds.), EIsevier Science, 1998,
865.
12. Isao Mochida and Kenjiro Takeshita,J Phys. Chem. 78,16 (1974) 1654.
13. J.H. Wang, Account Chem. Res. 3 (1970) 90.
14. Jubaraj B. Baruah and Lipika Nath, Ind J Chem. 36A (1997) 795.
15. R.G. Wilkins, The Study of the Kinetics and Mechanisn of Reactions of
Transition Metal complexes, Allyn and Bacon, Boston, 1974.
16. Susanta K. Sengupta and Preeti Lahiri, Ind J Technol. 30 (1992) 172.
17. O. Spalek, J. Balez, and I. Ponseka, J Chem. Soc. Faraday Trans. 1, 78
(1982) 2349.
193
Chapter 8
Hydroxylation of phenol to hydroquinone using zeolite
encapsulated complexes
Abstract
Hydroxylation of phenol is a reaction that has great commercial importance. Phenol hydroxylation
was carried out at room temperature using H
20:!
as oxidant. Encapsulated Cu(II) complexes of
morpholine-V-carbodithioate and salicylaldehydesemicarbazone were found to be very effective
for the selective oxidation of phenol. The reaction products were analyzed using gas
chromatograph. Due to size and shape selective effects, hydroquinone was obtained as the oniy
product. Influence of various parameters like reaction time, amount of catalyst, oxidant/substrate
ratio, solvent, and temperature were studied. Poison resistance and recyclability were also
checked. The complexes get deactivated by the addition of poison, and exhibit some loss in
activity on recycling.
8.1. Introduction
Organic synthesis today emphasizes on achieving selectivity of products. Weisz
and coworkers were among the first to report shape selective catalysis using
zeolite [1-3]. According to Weisz, the unique role and purpose of a catalyst is to
provide selectivity to direct chemical transformation along a specific, desired path.
Since these pioneering studies, shape selective catalysis has made great strides,
and has led to the development of many new industrial processes.
The catalytic activity shown by the encapsulated complexes can be
correlated to the well-defined structure of the active site inside the zeolite pore.
Now attempts are being made by researchers to develop new zeolite compounds
that mimic biological enzymes. These so called "zeozymes" are constructed by
trapping various metal complexes inside the crystalline framework of zeolites [4].
The unique feature of these zeozymes is that the catalyst remains in the solid phase
during the entire reaction and can be easily filtered off at the end of the reaction.
This aspect is significant, as it would reduce processing cost during large-scale
application.
In the present study, zeolite encapsulated complexes of Sf'O, SPP, SALSC,
SB, DTC and DMG have been used for the oxidation of phenol with hydrogen
peroxide with the main aim of obtaining the p-isomer as the only product. The
results of these studies are presented in this chapter.
8.2. Experimental
8.2.1. Materials
The details of the materials used for the catalytic study are given in Chapter
2. Zeolite encapsulated complexes of sro, SPP, SALSC, SB, DTC and DMG
were used for catalytic studies. Their neat analogues were also prepared and used
195
In catalytic studies to enable a comparison between neat and encapsulated
complexes.
8.2.2. Procedure for the hydroxylation of phenol
The hydroxyl ation of phenol was carried out in a 50 ml round bottomed
flask fitted with a reflux condenser. Phenol, catalyst, and water were introduced
into the reaction vessel and the reaction was initiated by adding hydrogen
peroxide. The mixture was stirred for a definite period of time using a magnetic
stirrer and a magnetic paddle at the set temperature. The reaction conditions are
given in the tables for each experiment. The catalyst was removed by filtration and
the analysis of the reaction product for hydroquinone was done using a gas
chromatograph. An SE-30 column was used for separating various components in
the reaction mixture. The components can be differentiated from the peaks that
appear on the recorder. The concentration of the component in the mixture was
determined from the peak area for that component.
The reaction was carried out under different conditions for testing the
influence of various parameters like time, amount of catalyst, oxidantlsubstrate
ratio, solvent and temperature. The recyclability of the catalyst and the poison
resistance were also studied.
8.3. Results
8.3.1. Screening studies
The data of the catalytic activity of SALSC, DTC, SPP, SPO, DMG and
SB encapsulated complexes in the hydroxylation of phenol using dilute hydrogen
peroxide as oxidant at room temperature is given in Table 8.1. A graph showing
percentage conversion of phenol to hydroquinone for the different
196
Table 8.1 Conversion of phenol to hydroquinone
% Conversion
SALSC MDTC SPP SPO DMG SB
FeY 0.2 0 0
CoY 3 4 1
NiY 1 2 1 0.5
CuY 97 99 2 0 0.5
PdY I 2 0.5
%Conversion to hydroquinone
"0
90
80
70
60
50
40
30
20
re
o .
F,Y CoY N,Y
Ligand
C,Y POY
IJ % Conversion SALSC
'0% ConYefSion MOTe
:a% Conversion see
:El % ConYefSion SPO
I % Conversion OMG
III % ConWlt"Sion S8
Figure 8.1 Conversion of phenol to hydroquinone
Reaction condition
Volume of phenol
Duration
Temperature
Weight of catalyst
Volume of H
2
0
2
Volume of solvent (H
2
0 )
: 1 ml
:4 h
:30QC
:40mg
:5 ml
:5 ml
197
complexes is shown in Figure 8.1. From the table it is clear that the maximum
conversion to hydroquinone was obtained in the case of CuYMDTC, with
CuYSALSC coming next. Water was used as a solvent in all the reactions.
8.3.2. Blank run
The hydroxylation of phenol was carried out using CuY under conditions
identical to those of the screening experiments. It was found that the only product
formed was catechol, and no hydroquinone was formed. Experiments were also
carried out using neat complexes of MDTC and SALSC [{Cu(MDTC)ll and
CuClSALSC]. Tarry products were obtained which could not be identified. The
reaction was repeated under identical conditions without adding catalyst. No
catechol or hydroquinone was found to be formed in this case.
8.3.3. Factors influencing the hydroxylation of phenol to form hydroquinone
CuYSALSC and CuYMDTC were found to be very effective for the
hydroxylation of phenol to hydroquinone, So further studies were made using
CuYSALSC and CuYMDTC and varying the following parameters to understand
the factors influencing the hydroxylation of phenol:
i) reaction time
ii) amount of catalyst
iii) HzO
z
- phenol ratio
iv) solvents
v) reaction temperature.
8.3.3.1. Influence of reaction time
The reaction was carried out for different durations varying from one hour
to six hours, using CuYMDTC and CuYSALSC, and the resulting products were
analyzed.
198
Table 8.2 Effect of reaction time on conversion of phenol to hydroquinone
Duration CuYSALSC CuYMDTC
(h)
% Conversion % Hydroquinone % Conversion % Hydroquinone
1 1 0 10 5
2 64 30 99 98
3 85 77 99 98
4 95 92 99 99
5 99 99 99 tar formed
100 , ' _______x
I
)<
/
80
/

c
0
60
c
/
"5 /
~
/
e
-c
40
>
I-x-CuYSALSC I
s:
15
if' /
! ---+-- CuYMDTC !
20
V 0
2 3 4 5
Time (h)
Figure 8. 2 Effect of reaction time on conversion of phenol to
hydroquinone
Reaction condition
Volume of phenol :1 ml
Volume ofH
202
:5 ml
Volume of solvent (H
20)
:5 ml
Weight of catalysts :50 mg
Temperature :30 QC
.'
" ,
j.)
i :::
\:.
.~
.,
-e-,
. ~ : .'"
199
The percentage conversion to hydroquinone is given in Table 7. 2, and the graph
showing the percentage of hydroquinone formed with time is presented in Figure
7.2. The percentage conversion of phenol to hydroquinone increased with time.
After one hour there was a sharp increase in the percentage conversion of phenol.
In the case of CuYMDTC, 99 % conversion with 100 % selectivity was
obtained within two hours. The reaction was continued for five hours and the
products were analyzed at the end of each hour. Tarry products were obtained at
the end of 5 hours. The darkening of the solution with time indicates tar
formation. CuYSALSC took 5 hours for 99 % conversion with 100 % selectivity.
Tarry products were formed after 5 hours. The optimum time for the reaction
using CuYMDTC was taken as 2 hours. and for the reaction using CuYSALSC it
was taken as 5 hours. After an optimum time of reaction, only tarry products are
formed.
8.3.3.2. Influence of the amount of catalyst
The progress of the reaction was studied by varying the amount of catalyst
from 20 mg to 50mg. The results are presented in Table 8.3, and Figure 8. 3 give
the percentage conversion of phenol to hydroquinone against varying amounts of
catalyst. It can be seen that the percentage conversion to hydroquinone increases
with increase in the amount of the catalyst. In the case of CuYMDTC. the
maximum conversion was obtained with 30 mg of catalyst while in the case of
CuYSALSC maximum conversion with 100 % selectivity was reached with 40
mg of catalyst. Further increase in the weight of the catalyst had no effect on the
oxidation of phenol to hydroquinone.
200
Table 8.3 Effect of the amount of the catalyst on the conversion of phenol to
hydroquinone
Weight of CuYSALSC CuYMDTC
catalyst
% % % % Hydroquinone
(mg)
Conversion Hydroquinone Conversion
20 4 0 50 55
30 33 31 99 99
40 99 99 99 99
50 99 99 99 99
100
/1
I!I
80
60
~ o
40
20
/ -D-CuYMDTC
0

20 25 30 35 40 45 so
weight of catalyst (mg)
Figure 8.3 Effect of the amount of the catalyst on the conversion of phenol to
hydroquinone
Reaction condition
Volume of phenol :1 ml
Volume ofH
202
:5 ml
Volume of solvent (H
20)
:5 ml
Temperature :30 QC
Duration
CuYSALSC : 5 h
CuYMDTC : 2 h
201
8.3.3.3. Influence of H,O,: phenol (oxidant/substrate) ratio
The reactions were carried out under identical conditions but by varying
the volume of hydrogen peroxide. The results showing the dependence of
percentage conversion to hydroquinone on the volume of H
2
0
2
are given in Table
8.4. The reaction conditions are also given in the table. From the table it is seen
that percentage conversion increases with increase in the volume of H
2
0
2
{keeping the volume of phenol constant (lm!.)} to a certain point, and thereafter
an increase in the volume ofH
2
0
2leads
to the production of tarry products.
Table 8.4 Effect of the amount of hydrogen peroxide on the conversion of phenol
to hydroquinone
Volume
ofH
2
0
2
CuYSALSC CuYMDTC
(ml)
4
5
6
% Conversion % Hydroquinone % Conversion % Hydroquinone
90 90 91 90
99 99 99 99
99 Tarry products 99 Tarry products
Reaction condition
Volume of phenol
Volume of solvent (H
2
0 )
Temperature
Weight of catalyst
Duration
CuYSALSC:
CuYMDTC:
:1 ml
:5 ml
:30 QC
:50mg
5 h
2h
202
8.3.3.4. Influence of solvents
The reaction was studied in different solvents like water, acetone,
methanol and benzene under similar reaction conditions. The effect of solvents on
the activity and selectivity of the catalysts is indicated in the Table 8.5.
Table 8.5 Effect of various solvents on the conversion of phenol to hydroquinone
CuYSALSC CuYMDTC
Solvents
% % % %
Conversion Hydroquinone Conversion Hydroquinone
Methanol 99 0 99 0
Benzene 99 0 99 0
Acetone 99 0 99 0
Water 99 99 99 99
Fonnalion of hydroquinone in different solvents
'00
"
.CuYSALSC
ElCuYMDTC
-------- -------
Figure 8.4 Effect of solvents on the conversion of phenol to hydroquinone
Reaction condition
Volume of phenol :1 ml
Volume ofH20z :5 ml
Volume of solvent (H20) :5 ml
203
Temperature
Weight of catalyst
Duration
CuYSALSC : 5 h
CuYMDTC : 2 h
:30 QC
:50 mg
Water was found to be the best solvent for phenol hydroxylation with 99
% conversion and 100 % selectivity towards hydroquinone. Tarry products were
formed in methanol and acetone. There was no reaction at all when benzene was
used as solvent.
8.3.3.5. Influence of reaction temperature
To study the influence of temperature on the conversion of phenol to
hydroquinone, the reaction was studied at three different temperatures, 30 QC,
50 QC, and 70 QC. The maximum conversion to hydroquinone was obtained at
room temperature (30
QC).
The yield of products in many reactions depends on
temperature. The results of the study are given in Table 8.6. With CuSALSC and
CuYMDTC, the percentage conversion to hydroquinone decreased to 0 % as
temperature was raised to 50 QC. At 70 QC, tarry products were formed with both
CuYSALSC and CuYMDTC.
Table 8.6 Effect of reaction temperature on the conversion of phenol to
hydroquinone
Temperature CuYSALSC CuYMDTC
('C)
% % % %
Conversion Hydroquinone Conversion Hydroquinone
30 99 99 99 99
50 99 0 99 0
70 99 0 99 0
204
Reaction condition
Volume of phenol
Volume ofH
2
0
z
Volume of solvent (H
2
0 )
Weight of catalyst
8.3.4. Addition of poison
:1 ml
:5 ml
:5 ml
:50mg
Duration:
CuYSALSC
CuYMDTC
:5 h
: 2 h
The poison resistance of the catalysts was studied by adding some foreign
compound to the catalyst reactant system. The effect of poison on CuYSALSC
and CuYMDTC was studied by examining the activity of the catalyst in presence
of traces of pyridine. The percentage loss in activity would indicate the extent of
deactivation on poisoning. This was carried out in two ways:
i) by directly adding pyridine to the catalyst, and
ii) by adding pyridine after adding phenol to the catalyst.
The catalysts were completely deactivated in the first case. In the second case the
two catalysts CuYSALSC and CuYMDTC were found to be resistant to the action
of poison.
8.3.5. Recycled catalyst
The reaction was carried out using used catalysts. The catalyst was filtered,
washed with acetone, and activated in the oven for two hours. This was to remove
any impurities attached to the active sites of the catalyst. X - Ray diffraction
studies of both the used catalysts (CuYMDTC and CuYSALSC) showed no loss
in crystallinity. The FTIR spectra of the used and unused catalysts were similar.
Catalytic runs were carried out under similar conditions using the recycled
catalyst. The results obtained for CuYSALSC and CuYMDTC are shown in
Figure 8.5, and the percentage conversions are given in Table 8.7. CuYSALSC
showed - 60 % activity compared to the activity with fresh sample while
CuYMDTC showed - 50%.
205
Table 8.7 Effect of recycled catalyst on the conversion of phenol to hydroquinone
Recycled
catalyst % Conversion % Hydroquinone % Loss in activity
CuYSALSC 60 60 40
CuYT\1DTC 55 51 49
CuYSALSC
49'V
'--- I I ~ ~ ~ J
Figure 8.5 Comparison ofthe activity and loss in activity of recycled catalyst
Reaction condition
Volume of phenol
Volume ofH
2
0
2
Volume of solvent (H20)
Weight of catalyst
Duration:
CuYSALSC
CuYT\1DTC
:1 m!
:5 ml
:5 ml
:50 mg
:5 h
: 2 h
206
8.4. Discussion
A comparison of the catalytic activity and selectivity of the zeolite
encapsulated SALSC, MOTC, SPO, SPP, DMG and SB complexes reveals that
CuYSALSC and CuYMDTC are the most efficient catalysts for the hydroxylation
of phenol to hydroquinone. There are reports that redox zeolites act as active
catalysts for hydrogen peroxide mediated oxidation reactions such as the
hydroxylation of phenol to the corresponding hydroxylated aromatics [3]. This
could be the result of various factors like surface area, pore volume, redox
properties of metal complexes, and the electric field gradient inside the zeolite.
Phenol hydroxylation reaction can take place at the extemal surface as well
as in the intemal pores. The reaction in the internal pores leads mainly to the
formation of hydroquinone whereas reaction on the external surface leads to the
production of catechol as well as tarry products. The formation of hydroquinone as
the only product suggests that the reaction occurs entirely inside the pores of the
zeolite. This might be due to the ligand field effect, which improves the selectivity,
and stability of the catalyst. Here the preferential formation of hydroquinone (para
isomer) is the result of product shape selectivity. Constraints imposed on the
diffusion of the products from the pores to the outside bring about enhanced para
selectivity. This is indicated in Figure 8.6.
From the percentage conversion taken as a function of the reaction time for
the two catalyst systems CuYSALSC and CuYMDTC, it is found that there is an
optimum time for maximum hydroquinone conversion. The appearance of an
induction period and the increase in the percentage conversion with time suggest
the involvement of a free radical mechanism. On carrying out the reaction for a
longer time, unwanted products were formed. The darkening of the solution with
time indicates tar formation.
207
+ H,o,
Figure 8.6 Influence of shape selectivity
The minimum amount of catalyst required for a particular reaction was
also optimized. There was no effect on increasing the weight of the catalyst above
a certain level.
The study of the effect of reaction temperature on hydroquinone
conversion shows that the maximum yield is obtained at room temperature. As
temperature increased to 50C, percentage conversion decreased. The reduction
in the percentage conversion to hydroquinone can be due to two reasons:
i) the formation of other products resulting from the slow diffusivity at
higher production ratio from the pore leading to further reaction;
ii) the decomposition of hydrogen peroxide at higher temperatures.
At 70C, tarry products are formed with both CuYSALSC and CuYMDTC. This
can be due to the exothermic nature of the reaction or due to over oxidation or due
to side reactions.
208
The study of the effect of solvents on the activity shows that water is the
best solvent for hydroquinone conversion. This is because phenol and hydrogen
peroxide can reach the active sites more easily with water than with organic
solvents. It has been reported that the activity for this reaction in water is much
higher than in any other solvent [5]. No hydroquinone conversion was observed
with methanol and acetone. This can be due to the radical scavenging property of
methanol and acetone, which removes the reactive phenol radicals. In benzene
there is no conversion at all, which might be due to the insolubility of hydrogen
peroxide in benzene.
The hydrogen peroxide phenol ratio is important in deciding the activity.
The maximum conversion to hydroquinone was obtained when the hydrogen
peroxide> phenol volume ratio was 5. Increase in the volume ofH202 beyond that
results in the production of tarry products. The formation of tarry products might
be due to over oxidation of phenol at high hydrogen peroxide concentration.
Both the catalysts were completely deactivated on adding pyridine. The
formation of catalytically inactive complex with pyridine with no vacant
coordination site might be the reason for the deactivation, which is indicated by
the colour change of the complex. However, CuYSALSC and CuYMDTC
showed no loss in activity on adding pyridine to the mixture of phenol and the
catalysts. This again suggests the possibility of the existence of phenoxide ions,
which get attached to the catalysts.
CuYSALSC and CuYMDTC can be reused after the used samples are
washed and heated (to activate the catalysts by removing the by-products
covering some active centres) [6]. The phenol conversion is in the range 50 - 60
% compared to that for the first run. This again confirms that the complex inside
the zeolite pore is stable. The reaction of metal complexes with peroxides is
important in organic synthesis since peroxides are useful forms of "active
209
oxygen" reagents for oxidative transformation [7]. Many possible mechanisms
have been suggested for phenol hydroxylation. As zeolite encapsulated complexes
are heterogenised homogeneous catalysts, the hydroxylation of phenol can be
assumed to follow a heterogeneous - homogeneous reaction mechanism [8-9].
The high susceptibility of the aromatic ring of phenols towards oxidation
can be due to the generation of the phenoxy radical via proton removal. The
generation of phenoxy radical may occur on the catalyst surface. There have been
reports of proton transfer from phenol to a potential active site prior to
hydroxylation during enzyme catalysis of phenol using phenol hydroxylase.
The mechanism involving homolytic hydroxyl radical pathway and
heterolytic pathway involving HjOt ion is presented in Scheme I and Scheme n
[10]. The first step in both the schemes involves the reaction between phenol and
the complex, in which an electron is transferred to the metal ion. This results in
the formation of phenoxy radical on the catalyst surface. At the same time the
catalytic surface can trigger the homolytic cleavage of hydrogen peroxide, or form
the very reactive species Hj 0
2
+, the acidic sites of zeolite providing the acidity
requirements for the formation of this species. The formation of catechol and
hydroquinone is believed to be the result of the attack of HO, on the benzene ring
or by the attack of Hj 0
2
",
As per the inferences from our study, the mechanism occurring in this case
can be the one following Scheme Il, utilizing n,o,'. Since the catalyst support is
zeolite there is strong possibility for the formation of H]Ot. The acidity
requirements for the formation of Hj02+ may be supplied by the acid sites.
This is further supported by the fact that hydroquinone conversion is maximum at
a very high concentration of hydrogen peroxide. The presence of acid sites of
zeolite and the high concentration of hydrogen peroxide shift the above
210
equilibrium to the right. This explains the high conversion observed with high
H
20Z
/phenol ratio.
+
Cu
2
+
6
+
0"
0 0
6

6

6
1OH"
1
OH"
0 0
6:"
6
H OH
1 1
;l:;OH
i..

OH
Scheme I Mechanism of phenol hydroxylation
211
OH
6 6
+
CU
2
+ +
W
+ CU
O 0 0
6
+
H
302
+
6:H
6
H OH
1
1
OH OH
oOH
Q

OH
Scheme 11 Mechanism of phenol hydroxylation
212
References
I. P.B. Weisz and V.J. Frilette, J Phys. Chem. 64 (1960) 382.
2. P.B. Weisz, V.J. Frilette , R.W. Maatman, and E.B. Mower, J Catal. 1
(1962) 307.
3. Chandra R Jacob, Saji P. Varkey, and Paul Ratnasamy, Appl. Cata!' A. 168
(1998) 353.
4. R.F. Parton, D.E. De Vos, and P.A. Jacobs in Proceedings of the NATO
Advanced Study Institute on Zeolite Microporous Solids: Synthesis,
Structure and Reactivity, E.G. Derouane, F. Lemos, C. Naccache, and F.R.
Ribeiro (Eds.), Kluwer, Dodrecht, 555, 578, 1992.
5. Dirk E. De Vos, Mieke Dama, Bert F. Sels, and P.A. Jacobs, Chem. Rev. 102
(2002) 3615.
6. Wei Zhao, Yunfei Luo, Peng Deng, and Quanzhi Li, Catal. Left. 73, 2-4
(2001) 199.
7. Roger A. Sheldon and Jay K. Kochi, Metal- Catalyzed Oxidations of
Organic Compounds, Academic Press, New York, 1981.
8. C. Meyer, G. Clement, and J.c. Balaceanu, Proc. Jrd Int. Congr. on
Catalysis, I (1965) 184.
9. A. Sadana and J.R. Katzer, J Catal. 35 (1974) 140.
10. G. Bellussi and C. Perego in Handbook ofHeterogeneous Catalysis, Vo!. 5,
G. Ertl, H. Knozinger, and J. Weitkamp (Eds.), Wiley-VCl-l, New York,
2329, 1997.
213
Chapter 9
Catalytic activity of zeolite encapsulated complexes in the
partial oxidation of benzyl alcohol to benzaldehyde
Abstract
Molecular sieves containing transition metal complexes exhibit very good properties as catalysts
for a variety of oxidation reactions with peroxide as oxidant. The partial oxidation of benzyl
alcohol using zeolite encapsulated complexes was carried out in the presence of hydrogen peroxide
as oxidant at room temperature. Benzaldehyde was the only detectable product. The influence of
various parameters such as temperature, amount of catalyst, and oxidant to substrate ratio was
studied. Significant variation was observed on changing the solvent. The effect of catalyst poison
was studied by adding pyridine. Higher activity was observed in the case of PdYDMG and
CuYSPP. The efficiency of the catalysts on recycling was also studied.
9.1. Introduction
The oxidation of primary and secondary alcohols to aldehydes, ketones and
carboxylic acids is a pivotal reaction in organic syntheses [1-2]. Traditionally such
reactions have been performed with inorganic oxidants containing chromium(vI)
reagents. The emphasis in recent times is on developing recyclable catalysts to
bring about oxidation reactions using clean oxidants such as oxygen and hydrogen
peroxide, under mild conditions compared to classical oxidations. Metal
complexes have attracted attention as dioxygen activating catalysts. Per-oxo
complexes act as oxidants in different reactions [3]. Per-oxo compound are
generated by adding hydrogen peroxide to metal complexes But these
homogeneous systems have several drawbacks, as they show a tendency for self
degradation, and there is difficulty in reusing them. One means of overcoming
these problems was to encapsulate them in zeolite pores, as zeolite encapsulated
systems have the advantages of both homogeneous and heterogeneous catalysts.
Zeolite encapsulated per-oxo compounds provide an attractive route for the
preparation of synthetic intermediates and other oxygen-containing organic
substrates. Partial oxidation of benzyl alcohol is of great importance in the
preparation of fine chemicals. In this chapter the results of our studies on the
oxidation of benzyl alcohol to benzaldehyde in the presence of PdYOMG and
CuYSPP complexes as catalysts are presented.
9.2. Experimental
9.2.1. Materials
Zeolite Y encapsulated complexes ofSPO, SPP, SALSC, SB, OMG, and
MOTC were prepared and characterized. Details of all the materials used for the
activity studies are explained in Chapter 2. Neat complexes of the catalytically
active Iigands were also synthesized.
215
9.2.2. Procedure for the partial oxidation of benzyl alcohol to benzaldehyde
Benzyl alcohol, hydrogen peroxide, the activated catalyst, and the solvent
were taken in a reaction flask of 50ml capacity. The reaction mixture was stirred
for a definite period of time using a magnetic paddle. The conditions specific to
each test are indicated along with the results in the table for each reaction.
After the reaction the solution was filtered to remove the catalyst. Then the
aqueous and organic layers were separated. The organic layer was washed with
water and analysed using thin layer chromatography (TLC) to detect the formation
of benzaldehyde. It was finally analysed using gas chromatography to quantify the
amount of substrate and product. The reaction was repeated, varying parameters
like time, amount of catalyst, oxidantlsubstrate ratio, solvent, and temperature. The
poison resistance and the activity on reusing the catalyst were examined here also.
9.3. Results
9.3.1. Screening studies
The partial oxidation of benzyl alcohol to benzaldehyde was studied using
SPO, SPP, MOTC, OMG, SB and SALSC complexes in the liquid phase. The
activities corresponding to these zeolite-encapsulated complexes are gIven In
Table 9.1, and the percentage conversion to benzaldehyde for the different
complexes is shown in Figure 9.l. Of all the complexes, PdYOMG and CuYSPP
showed more than 95% conversion with 100% selectivity, while PdYSALSC
showed only 49% conversion. Conversion is represented as turnover number
(TON) calculated as moles of benzyl alcohol converted per mole of metal ion
present in the catalyst.
216
Table 9.1 Activity for benzyl alcohol oxidation
DMG
FeY
CoY
NiY
CuY
PdY
99
100
90
80
c
70 0
r! 60
CD
50 >
c
0
40
u

30
0
20
10
0
Percentage of benzaldehyde
SPP SPO SALSC MDTC
0.1 0.1 0.1
4.0 0.5
0.5
6.0
0.1 0.1
0.1
0.1
98 6.2
3.5
25
49
2.1
Conversion of benzyl alcohol to benzaldehyde
'-
OMG
,... ....
I511III:
: .. &lilt:[
&l1li::: 1iIIIIO[]II'
SPP SPO SALSC MOTC SB
I
Percentage of Benzaldehyde
Ligand
SB
4.0
leFeY!
I
_CoY!
,1iI NiY !
i
OCUY
,
!_PdY: I
I
,
Figure 9.1 % Conversion of benzyl alcohol to benzaldehyde
Reaction condition
Oxidant to substrate mole ratio: 1.5
Temperature : 30 QC
Catalyst weight : 60 mg
Solvent : water
Duration : 4 hours
217
9.3.2. Blank run
The oxidation of benzyl alcohol to benzaldehyde was carried out under
conditions identical to those of the screening experiments but without adding the
catalyst. Benzaldehyde was not formed. This makes it clear that hydrogen peroxide
by itself is unable to oxidize the benzyl alcohol in the absence of the catalyst. The
reactions were carried out using NaY, Cu Y and PdY. Formation of benzaldehyde
was not observed in these cases either. Experiments using neat complexes ofDMG
and SPP were also carried out. The absence of benzaldehyde in all these cases
indicated that the conversion was possible only in presence ofthe catalysts.
9.3.3. Factors influencing the partial oxidation of benzyl alcohol to
benzaldehyde
The oxidation of benzyl alcohol to benzaldehyde was studied under varied
conditions, and the results observed were tabulated. It is seen that PdYDMG and
Cu YSPP give good conversion for benzyl alcohol oxidation. Further studies were
done using PdYDMG and CuYSPP by varying the following parameters:
i) reaction time
ii) amount of catalyst
iii) oxidant to substrate ratio
iv) solvents
v) reaction temperature.
9.3.3.1. Effect of reaction time
The effect of reaction time was studied by carrying out the reaction for
different time intervals ranging from 1 hour to 5 hours using PdYDMG and
Cu YSPP. The percentage conversion increased with increase in time. At the end of
four hours 99% conversion was obtained with both the catalysts, with 100 %
218
selectivity to benzaldehyde. An induction time of 2 hours was observed in these
cases.
Table 9.2 Effect of reaction time on conversion of benzyl alcohol
Time PdYDMG CuYSPP
(h) % Conversion % Benzaldehyde % Conversion % Benzaldehyde
0
2 0
3 40
4 99
5 99
2
0
0
36
99
99
,
Time (h)
0
0
34
99
99
,
5
0
0
33
99
99
!OPdYDMG
ill cuYSPP
- --------- -------------- - -
Figure 9.2 Reaction time versus % conversion
Reaction condition
Oxidantfsubstrate mole ratio
Weight of catalyst
Solvent
Temperature
: 1.5
: 60mg
: water
: 30C
219
There was no change in the amount or product on carrying out the reaction
for 5 hours. The effect of contact time on conversion and selectivity is represented
in Table 9.2 and this is shown graphically in Figure 9.2.
9.3.3.2. Effect of the amount of catalyst
The reactions were carried out with PdYDMG and CuYSPP under identical
conditions by varying the amount of catalyst from 20mg to 60mg. It was found
that activity increased with increase in the amount of the catalyst. The results are
tabulated in Table 9.3 and are presented graphically in Figure 9.3. The maximum
conversion was obtained with 60mg of catalyst. So for all the reactions the amount
of the catalyst was fixed as 60mg.
100
80
c
60
:u
>
c
40

20
o
20 30 40 50 60
Weighl ot calslysl
Figure 9.3 Effect of the amount of catalyst on benzyl alcohol oxidation
Reaction condition
Oxidantlsubstrate mole ratio : 1.5
Solvent : water
Temperature : 30 QC
Duration : 4 hrs
220
Table 9.3 Effect of the amount of catalyst on benzyl alcohol oxidation
Amount of PdYDMG CuYSPP
catalyst % % % %
Conversion Benzaldehyde Conversion Benzaldehyde
20 0.1 0
40 10 15 5 10
60 99 99 99 99
9.3.3.3. Effect of oxidant: substrate ratio
The effect of oxidant: substrate ratio on benzaldehyde oxidation was studied in the
case of PdYDMG and CuYSPP. The extent of oxidation depends on the
concentration of oxidant in the reaction mixture. The reaction was repeated under
identical conditions but by varying the oxidant: substrate ratio. The percentage
conversion to benzaldehyde increased as oxidantlsubstrate mole ratio increased.
The maximum conversion was obtained at oxidantlsubstrate mole ratio 1.5. There
was no further change in conversion as the ratio was increased. So further studies
were carried out fixing the oxidantlsubstrate mole ratio at 1.5. The results are
presented in Table 9.4.
Reaction condition
Temperature
Duration
Weight of catalyst
Solvent
: 30 QC
: 4 hrs
: 60mg
: water
221
Table 9.4 Effect of varying the oxidantlsubstrate ratio on the conversion of phenol
to hydroquinone
PdYDMG CuYSPP
Solvent
% % % %
Conversion Benzaldehyde Conversion Benzaldehyde
2 0 0 0
1.5 99 99 99 99
2.5 99 99 99 99
9.3.3.4. Effect of solvents
The study of the oxidation of benzyl alcohol was carried out in two solvents.
benzene and water. The effect of solvents was studied for two catalysts. PdYDMG
and CuYSPP. The results are presented in Table 9.5 and represented pictorially in
Figure 9.4. Using water as solvent, 99% conversion was observed with 100%
selectivity to benzaldehyde. In benzene, 100% conversion of benzyl alcohol
occurred but no benzaldehyde was formed.
I

Figure 9.4 Effect of solvents on the conversion of benzyl alcohol to benzaldehyde
222
Table 9.5 Effect of solvent on benzyl alcohol oxidation
PdYDMG CuYSPP
Solvent % % % %
Conversion Benzaldehyde Conversion Benzaldehyde
Water
Benzene
Reaction condition
99
99
Oxidantlsubstrate mole ratio: 1.5
Temperature : 30C
Duration : 4 hrs
Weight of catalyst : 60 mg
98
o
9.3.3.5. Effect of reaction temperature
99
99
98
o
The oxidation of benzyl alcohol was carried out at three different
temperatures 30, 50
0
and 70C, using PdYDMG and CuYSPP to find the
influence of temperature on the partial oxidation reaction. The results are given in
Table 9.6 and the variation in activity with temperature is shown in Figure 9.5. All
the reactions were carried out under identical conditions except the difference in
temperature. On increasing the temperature the activity was found to decrease in
the case of both the catalysts.
Reaction condition
Oxidantlsubstrate mole ratio
Solvent
Weight of catalyst
Duration
: 1.5
: water
: 60mg
: 4 hrs
223
Table 9.6 Effect of reaction temperature on benzyl alcohol oxidation
Reaction PdYDMG CuYSPP
temperature % Conversion % Benzaldehyde % Conversion % Benzaldehyde
30C
99
50C
2
70C
100
80
c
0
60
i
40 u
.,.
20
0
98
1.2
0.6
..,...
---
30
D
e
50
D
e
Temperature
, PdYDMG
cuvspp
---
70
D
e
99
2
Cuvspp
PdYDMG
Figure 9.5 Effect of reaction temperature on benzyl alcohol oxidation
9.3.4. Addition of poison
98
1.5
0.1
The poison resistance ofPdYDMG and CuYSPP was evaluated by using pyridine.
Pyridine was added to the catalysts for poisoning them. Then the reaction was
carried out using the poisoned catalysts under identical conditions. Both the
catalysts were seen to be deactivated.
224
9.3.5. Recycled catalyst
Use of recycled catalyst is very important from the point of view of economical
considerations. The used catalyst was filtered off from the reaction mixture and
washed with acetone and dried in the oven at 120 QC for 2 hours. XRD patterns
were found to remain similar after the first run, indicating the retention of the
zeolite structure. IR spectra indicated that the complex was intact in the cages.
Pd(u) and CU(II) ions were not detected in the reaction phase of the mixture.
Therefore the Pd and Cu complexes have not leached from the zeolite pore during
the reaction. PdYDMG showed 67% activity after recycling while in the case of
Cu YSPP, there was no loss in activity on recycling. A comparison of the activities
of fresh and recycled catalysts is given in Figure 9.6 and the corresponding
conversions and percentage loss in activity are given in Table 9.7.
Table 9.7 Effect of recycled catalyst on benzyl alcohol oxidation
Catalyst % Conversion
used Before recycling After recycling % Loss in activity
PdYDMG
CuYSpp
Reaction condition
98
98
Oxidant! substrate mole ratio: 1.5
Temperature : 30 cC
Duration : 4 hrs
Solvent : water
Weight of catalyst : 60 mg
67
98
33
o
225
9.4. Discussion

PdYDMG CuYSpp
Loss in activity
[] Retained
Figure 9.6 Activity on using recycled catalysts
Zeolite encapsulated complexes of SPO, SPP, MOTC, SS, OMG and
SALSC were used in tne oxidation of benzyl alcohol. A comparison of the activity
indicates that PdYDMG and CuYSPP are the most effective catalysts for the
partial oxidation of benzyl alcohol, and there was 100% selectivity to
benzaldehyde. In terms of turn over number. the conversion for PdYDMG was
2540 and that for CuYSPP was 1682. PdYSALSC and CuYMOTC showed
moderate activity compared to PdYDMG and CuYSPP. Benzoic acid or any other
oxidation products were not detected in the reaction mixture on analysis using gas
chromatography. Thi s shows that zeolite encapsulated complexes are very
effective catalysts for the partial oxidation of benzyl alcohol, and that PdYDMG
and CuYSPP are eco-friendly catalysts in that they leave no waste products.
Similar behaviour of encapsulated complexes in the partial oxidation of benzyl
alcohol has been reported by Zsigmond [4] . The aqueous and the organic phases
were separated and the aqueous phase was tested for the presence of any metal
ions using atomic absorption spectroscopy after the screening studies. The absence
of metal ions in the solution phase indicated that no leaching of complexes had
226
occurred during the reaction. This shows that the complexes are intact inside the
zeolite pores. This explains the high stability of the complex inside the zeolite
pore.
The activities of the zeolite encapsulated catalysts were compared with
those of NaY, CuY, and PdY in partial oxidation reaction. The absence of the
oxidation product benzaldehyde with NaY indicates that the zeolite support in
itself does not ensure oxidation. Negative results with Cu Y and PdY indicate that
Cu
2
+ and Pd
2
+ ions by themselves are not able to influence benzyl alcohol
oxidation. Neat complexes of Pd(JI) and CU(JI) did not yield benzaldehyde as the
product on the oxidation of benzyl alcohol. This could be due to multi molecular
deactivation. On the other hand, zeolite encapsulated complexes showed high
activity. This can be explained by the site isolation theory. The complexes are held
within the zeolite pores by physical force. Confinement of the transition metal
complexes inside the cavities can activate the catalyst because of the strong
electrostatic interaction between the catalyst and the support. They are free to
move inside the pore but are prevented from coming out due to the restrictive size
of the pore openings. This prevents deactivation by multi molecular association. It
has been reported that a metal complex in the zeolite cage behaves like neat
complexes in solution [4].
The effect of varying the amount of the catalysts was studied. The
conversion increased with an increase in the amount of the catalyst upto a certain
limit at which maximum conversion was obtained, and further addition of catalyst
had no effect. The increase in conversion can be due to the increase in the number
of active sites.
In the case of both the catalysts, there was an induction period of 2 hours
after which the conversion to benzaldehyde slowly increased. The time required
for the completion of the reaction was 4 hours. There are reports that the oxidation
227
of benzyl alcohol is a slow process [5]. No further reaction products were observed
on continuing the reaction beyond 4 hours.
The choice of solvent was found to have a big influence on activity. Very
high activity was observed when the solvent used was water. Benzaldehyde
formation was not detected on using benzene as solvent. This can be due to the
low solubility of hydrogen peroxide in benzene.
The effect of temperature was studied, conducting the reaction at three
different temperatures, 30C, 50 C, and 70C. Conversion was maximum at room
temperature, and room temperature was taken as the optimum temperature for
benzyl alcohol oxidation. According to earlier reports, the reaction is to take place
at higher temperatures [6). In the present case, the conversion to benzaldehyde
decreased as the temperature was raised. This could be due to the decomposition
of hydrogen peroxide at higher temperatures. There is an induction period of 2
hours for this reaction. The decomposition of hydrogen peroxide must have
occurred within that time at higher temperatures.
The maximum conversion was obtained when the oxidantlsubstrate mole
ratio was 1.5. There have been reports of experiments in which the reaction was
carried out with the oxidantlsubstrate mole ratio at 2.5 [6). On adding poison to the
catalyst it was completely deactivated by the action of poison. One of the most
attractive features of zeolite encapsulated complexes is their ability to be recycled.
CuYSPP maintained the same activity after recycling, while PdYDMG showed
63% activity as compared to the fresh catalyst. This shows their tendency to
withstand the conditions of regeneration processes. The use of recycled catalyst
reduces the possibility of environmental pollution and enables the continuity of the
process.
According to the mechanism proposed by Puzari et aI., some excess
hydrogen peroxide is needed because the reaction involves the decomposition of
228
hydrogen peroxide. As per this mechanism, the catalytic oxidation of benzyl
alcohol by hydrogen peroxide involves two simultaneous processes:
decomposition of hydrogen peroxide itself resulting in the liberation of oxygen,
and the liberated oxygen reacting with benzyl alcohol [7]. In our observation, the
oxidation of benzyl alcohol takes place along a single route. Besides, the
oxidantlsubstrate mole ratio used in our method is lower than that reported earlier
[6]. The mechanism proposed here involves the interaction of the encapsulated
complex with hydrogen peroxide to form a catalyst precursor. The precursor is
formed by the coordination of an oxygen atom to the vacant metal site in the
complex. The catalyst precursor interacts with benzyl alcohol oxidizing it to
benzaldehyde. Similar mechanisms have been reported earlier [8-10].
Catalyst precursor
Catalyst
Scheme I Oxidation of benzyl alcohol to benzaldehyde
PdYDMG and CuYSPP, which were found effective in the oxidation of
benzyl alcohol, have square planar geometry. So there is enough space around the
metal for the coordination of one or more ligand. The precursor is formed by the
coordination of an oxygen atom to the vacant metal site in the complex.
The use of encapsulated systems has certain advantages over the use of
homogeneous systems. It is very easy to separate the encapsulated catalyst from
the reaction medium by simple filtration. Even the used catalyst showed high
activity while the free complex had no activity. Further, there is no need to use an
extra axial ligand as in the case of the homogeneous catalyst [4]. The other
advantages were the possibility of recycling and amenability to continuous
processing.
229
References
1. M. Hudlicky, Oxidations in Organic Chemistry, American Chemical
Society, Washington DC, 1990.
2. R.A. Sheldon, I.W.C.E. Arends, Gerd-Jan T. Brink, and A. Duksman, Acc.
Chem. Res. 35 (2002) 774.
3. Won K. Seok, Bull. Korean Chem. Soc. 20,4 (1999) 395.
4. A. Zsigmond, F. Notheisz, Z. Frater, and J. E. Backvall, in Heterogeneous
Catalysis and Fine Chemicals IV, H.U. Blaser, A. Baiker, and R. Pr ins
(Eds.) Elsevier Science B.Y. (1997) 453.
5. L. Palombi, L. Arista, A. Lattanzi, F. Bonadies, and A. Scettri,
Tetrahedron Lett. 37,43 (1996) 7849.
6. K.O. Xavier, J. Chacko, and K.K. Mohammed Yusuff, J Mol. Catal. 178,
(2002) 275.
7. A. Puzari, L. Nath, and Jubaraj B. Baruah, Ind. J Chem. 37 A (1998) 723.
8. R. A. Sheldon and J.K. Kochi, Metal Cata(vzed Oxidation of Organic
Compounds, Academic Press, New York, 1981.
9. Y. Kamiya and M. Kashima, Bull. Chem. Soc. Jpn. 46 (1974) 905.
10. A. Onopchencko and J.G.D. Schultz, J Org. Chem. 37 (1972) 2564.
230
Summary and conclusion
With increasing awareness about the dangers of environmental degradation,
research in chemistry is getting increasingly geared to the development of "green
chemistry," by designing environmentally friendly products and processes that
bring down the generation and use of hazardous substances. The priority in many
chemical processes now is to develop new solid catalysts with selective and
effective transformation possibilities for oxidation reactions with high yield.
Catalysis by metal complexes encapsulated in the cavities of zeolites and
other molecular sieves has many features of homogeneous, heterogeneous. and
enzymatic catalysis. Serious attempts have been made to gain product selectivity
in catalysis by means of enzyme mimicking. Immobilization of the active center in
the zeolite pores is the essence of synthetic biomimetic chemistry, and is
responsible for the shape selective properties of such catalysts. The notable
superiority of this kind of catalytic system is in single-site reactions, high
selectivity, and tenability (which are the advantages of homogeneous catalysts)
together with ease of separation of products from reactants, durability, stability. and
easy handling (advantages of heterogeneous catalysts). The catalytic activity
shown by the encapsulated complexes can be correlated to the structure of the
active site inside the zeolite pore.
Chapter 1 of the thesis presents a general introduction about the zeolite
encapsulated metal complexes and the catalytic activity of metal complex catalysts
in certain oxidation reactions, particularly in the decomposition of hydrogen
peroxide, hydroxylation of phenol, and the partial oxidation of benzyl alcohol.
Information regarding the methods used for the synthesis of zeolite encapsulated
complexes; the role of zeolites as micro reactors and as enzyme mimics, and the
interaction between the guest and the host is also included in this chapter. The
chapter ends with a discussion of the significance of the work undertaken and its
scope, followed by relevant references.
Chapter 2 presents the details regarding materials and methods used for
the synthesis and characterization of the complexes. The techniques employed for
the characterization of the complexes include CHN analysis, AAS, XRD, SEM,
BET surface area and pore volume, magnetic measurements, DRS, FTIR, EPR,
and TG. The methods and techniques used for carrying out catalytic activity
studies and estimating the percentage conversion are also included in this chapter.
Chapter 3 is divided into two sections: Section A and Section B. Section
A describes the synthesis and characterization of zeolite Y encapsulated Fe(III),
Co(n), Ni(n), and Cu(n) complexes of the Schiff bases NN'-bis(3-pyridylidene)-
1,2-phenylenediamine (SPO) and N,N' -bis(3-pyridylidene )-1 ,4-phenylenediamine
(SPP). Section B deals with Co(n), Ni(n), and Cu(n) complexes of salicylaldehyde
semicarbazone (SALSC).
Ion exchanged zeolites MY (M= Fe, Co, Ni, Cu, and Pd) were prepared
and characterized. All the encapsulated complexes were synthesized using the
flexible ligand method and purified by soxhlet extraction. SEM taken before and
after encapsulation indicates the absence of surface adsorbed species. XRD
patterns indicate that there is no loss in crystallinity for the zeolite structure even
after encapsulation. The Si/AI ratio of the encapsulated complex is the same as that
of the ion exchanged zeolite showing the retention of zeolite structure after
encapsulation. The lower surface area and pore volume of the encapsulated
complex compared to the ion-exchanged zeolite suggest encapsulation of the metal
complex inside the zeolite pore.
The coordination sites of ligands to the metals were arrived at from the IR
spectra. The strong bandsat 1581cm-
1
and 1610 cm-I in the spectra of the ligand
SPO and SPP are shifted to the lower energy region in the spectra of complexes,
232
indicating coordination of azo-methine nitrogen. The ligand absorption of SALSC
at 1620cm-
1
due to uc=o is shifted to a lower region (l550-1545cm-
l
) in the
complexes. On the basis of magnetic moment, electronic spectra, and EPR of
Cu(II) complexes, an octahedral geometry was tentatively assigned for the
encapsulated complexes of FeYSPO, CoYSPO, NiYSPO, FeYSPP, CoYSPP,
NiYSPP, CoYSALSC, and NiYSALSC; and square planar geometry was assigned
for CuYSPO, CuYSPP, and CuYSALSC from EPR spectra. The a
2
values between
0.87 and 0.91 for the copper complexes of SPO, SPP, and SALSC suggest an ionic
environment for Cu(u) ions. Neat SPO complexes of Co(u) and Ni(n) have been
reported to be square planar. The change in geometry in the case of Ni(n)
complexes can be attributed to the interaction of the oxide ions of the zeolite
framework. Only a qualitative idea regarding the decomposition patterns of zeolite
encapsulated complexes was obtained from the TO curves. The mass loss around
lOO-125C is due to loss of free water and coordinated water. The major mass loss
from 125-500 C can be due to the decomposition of the complexes. However
distinct stages of decomposition were not observed in any of the TO curves.
Chapter 4 presents a comparative study of neat and encapsulated 2-
styrylbenzimidazole complexes of Cu(u). The analytical data confirm the
composition of the neat complex to be Cu(SBhCh. The Si/AI ratio and the XRD
patterns of the encapsulated complex suggest that there is no collapse of the zeolite
framework on encapsulation. The formation of the complex in the zeolite cage is
further confirmed by surface area analysis, pore volume analysis, and the absence
of surface species was noted from SEM.
The IR spectral data indicate that the coordination of the ligand is through
the vC=N of styrylbenzimidazole. Magnetic moment, diffuse reflectance spectra
and EPR spectra of the two complexes suggest square planar structure for the neat
and encapsulated complexes. The TOIDTO of the neat complex gives clear idea
regarding the decomposition pattern of the complex whereas the TO of the
encapsulated complex gives only qualitative idea regarding the decomposition.
233
There are two stages of decomposition for the neat complex. The mass loss in the
first stage might be due to expulsion of one styrylbenzimidazole and the mass loss
in the second stage of decomposition corresponds to the removal of one
styrylbenzimidazole and two chlorine atoms. The TG curve of the encapsulated
complex shows that most of the intrazeolite free water molecules are released in
the temperature range 30-100 DC. The mass loss in the temperature range 100-534
DC can be due to the removal of ligands and coordinated water if any.
Chapter 5 is divided into two sections: Section A and Section B. Section
A deals with the synthesis and characterization of zeolite encapsulated complexes
of Fe(m) dithicarbamates of MDTC, EPDTC and IPDTC, and Section B deals
with the studies on CoOII), Ni(II), and Cu(n) morpholinedithiocarbamtes. The
SilAI ratio and the XRD patterns indicate the retention of zeolite framework.
Decrease in surface area and pore volume confirms encapsulation of complexes
and SEM indicates the absence of surface species. The positive shift of VN=CS
2
seen in the IR spectra is an indication of the interaction of the chelated sulphur
atoms with the metal. This indicates that coordination has occurred through
sulphur atoms.
Magnetic moment studies indicate octahedral structure for the encapsulated
Fe(m) complexes, suggesting that there is no change in structure on encapsulation.
The cobalt complex was found to be diamagnetic,suggesting a low spin octahedral
structure for the complex. The neat Ni(lI) dithiocarbamate complexes which were
reported to have square planar appears to have octahedral structure when inside
the zeolite cages. The Cu(u) complex which is square planar showed no change in
geometry on encapsulation. The structure of copper complex is further supported
by diffuse reflectance spectroscopy and the EPR spectrum. The ratio gill All is 109.2
cm for CuYMTDC suggesting a square planar environment around Cu (11) ion. The
value of a
2
is 0.9495, suggesting ionic environment for the Cu
2
+ ions in the zeolite
cavity. TG curves are continuous and so the stages of decomposition are not
distinct.
234
Chapter 6 reports the study of the zeolite encapsulated palladium
complexes of OMG, SALSC and MDTC. Chemical analysis reveals that the
complexes are present as monomeric species in the zeolite lattice. The lowering of
the surface area and pore volume values confirm the presence of the complex
inside the zeolite pore, and the XRO patterns indicate the retention of zeolite
crystallinity. The shift of the C=N stretching vibration to lower energy region of
ligand OMG; the positive shift of VN=CS
2
of morpholinecarbodithioate; and the
shift of vc=o stretch of salicylaldehyde semicarbazone to 1600 cm-I are observed
in the IR spectra of PdYOMG, PdYMOTC, and PdYSALSC respectively, which
support the coordination of the ligand to the metal in the complexes. All the
palladium-encapsulated complexes were found to be diamagnetic. Electronic
spectra also support square planar geometry for the complexes.
Chapter 7 deals with the studies on the decomposition of hydrogen
peroxide in the presence of complexes of sPa, SPP, SALSC, SB, MOTC, and
OMG. The catalytic activity of PdYSALSC, CuYMDTC and CuYSB was studied
in detail. PdYSALSC was found to be the most active catalyst in the
decomposition; CuYMOTC showed moderate activity, while CuYSB showed
much lower activity. The difference in activity can be correlated to the structural
features (geometry) of the complex. All the active complexes have square planar
geometry. Complexes having two free sites on the metal ion facilitates the reaction
of the intermediate LM(HOO) (n-I)+ with a second peroxide ion. The highly
reactive species formed enhances the decomposition process. The activity of the
encapsulated complex was much higher compared to that of the ion exchanged
zeolites indicating the influence of the ligand. In the present study, activity with
regard to the ligand was found to vary as follows:
SALSC > MOTC > SB.
Therefore it appears that the chelate fonnation and the ring size of the chelate have
a role in the activity.
235
The rate of decomposition reveals that the encapsulated complexes are
more active than the neat complexes. The decomposition of hydrogen peroxide
with each of the three catalysts was found to increase with increase in the weight
of the catalyst. A change in the solvent polarity of the reaction medium slowed
down the decomposition reaction indicating the involvement of the HOO- species
in catalysis. The direct addition of pyridine to the catalyst deactivated the
decomposition processes while the addition of pyridine to the mixture of hydrogen
peroxide and catalyst increased the rate of the reaction. This again confirms that
hydrogen peroxide decomposes according to the equation
H
2
0
2
--+ W + HOO-
HOO- being the species involved in catalysis. The activity of the Cu complexes
was retained / slightly lost on recycling while the activity of the encapsulated Pd
complex was considerably reduced.
Chapter 8 of the thesis presents our studies on the hydroxylation of phenol
using the encapsulated complexes. The hydroxylation of phenol to hydroquinone
was carried out at room temperature with 30 % hydrogen peroxide as oxidant.
From the screening studies, CuYSALSC and CuYMDTC were found to be the
most active catalysts. GC analysis of the product of the reaction revealed that these
catalysts exhibit high para selectivity in the product stream. Selectivity to
hydroquinone was greater than 95 % with both the catalysts. Apart from high
selectivity, high hydroquinone yield was also observed in both the cases. This
method is simple and economical and minimizes the formation of undesired
products. The reaction was screened using metal exchanged zeolite (CuY) and neat
complexes. On screening with CuY, catechol was found to be the only product. In
the case of neat complexes, tarry products were obtained, which might be due to
the presence of too many active centers. Thus encapsulation increases the activity
and para selectivity in these cases. Another feature of the reaction is that the
catalyst remains in the solid phase during the entire reaction and can be easily
filtered off at the end of the reaction. Activity increased with increase in the
amount of the catalyst up to a certain level. Further increase in the amount of the
236
catalyst had no effect at all. It was observed that there is an optimum time for each
catalyst for the completion of the reaction. The optimum time for CuYSALSC was
5 hours and that for Cu YMDTC was 2 hours. Unwanted products were obtained
when the reaction was allowed to proceed beyond this time. This might be due to
over oxidation or due to the decomposition of hydrogen peroxide. The induction
period observed in the cases of the two catalysts suggests a free radical mechanism
for these two catalysts.
The catalytic activity showed a strong dependence on the hydrogen
peroxide - phenol volume ratio. Tarry products were obtained on increasing the
volume of hydrogen peroxide. This can be due to the over oxidation of phenol at
high hydrogen peroxide concentration. The maximum yield of hydroquinone was
obtained when the H
2
02 _ phenol volume ratio was 5. This suggests that the
possible mechanism can be via the formation of H
3
0
2
+.
The yield of hydroquinone decreased as temperature increased from 30C
to 70 QC. The reduction in the percentage conversion to hydroquinone on
increasing the temperature to 50C may be due to the formation of some other
products resulting from the slow diffusivity at higher production ratio from the
pore, leading to further reaction or due to the decomposition of hydrogen peroxide
at higher temperatures. Tarry products are formed at 70C with both CuYSALSC
and Cu YMDTC, and this can be the result of the exothermic nature of the reaction
or due to over oxidation, or some other side reaction.
Water appears to be the best solvent for the reaction compared to methanol,
acetone and benzene. The radical scavenging methanol and acetone may remove
the reactive phenol radicals and affect the catalytic reaction. In benzene there is no
conversion at all. This can be because hydrogen peroxide is insoluble in benzene.
The catalyst showed good resistance to deactivation on adding poison (pyridine).
On recycling they showed 50-60% activity compared to unused catalysts. This
again confirms the stability of the catalysts inside the zeolite pore.
237
Chapter 9 deals with our studies on the partial oxidation of benzyl alcohol
to benzaldehyde. The oxidation was carried out using hydrogen peroxide as
oxidant in presence of PdYDMG and CuYSPP as catalysts. The product
(benzaldehyde) was detected using TLC and confirmed using Gc.
The catalytic activity of the complexes was tested for oxidation under
various conditions. The operating conditions like the amount of the catalyst,
reaction time, oxidant to substrate ratio, reaction temperature, and solvents have
been optimized. No further oxidation products were obtained on continuing the
reaction for four hours beyond the optimum time. Maximum conversion was
obtained at room temperature and the percentage conversion decreased with
increase in temperature. Activity was found to be dependent on the solvent used.
The poison resistance of the catalysts was checked by adding pyridine.
Both the catalysts were deactivated by poison. The activity of the recycled
catalysts was also studied. Cu YSPP showed no change in activity on first
recycling while the activity of PdYDMG was reduced to almost 67 % of the fresh
catalyst.
Both PdYDMG and Cu YSPP have square planar geometry indicating the
influence of structure in catalytic activity. It appears that the precursor formed by
the coordination of an oxygen atom to the vacant metal site in the complex might
be responsible for the oxidation of benzyl alcohol to benzaldehyde. Further more,
in these reactions, there was 100% selectivity towards benzaldehyde. In effect
these catalysts satisfy the basic need of green chemistry, making chemical
synthesis more environmentally friendly.
238

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy